arxiv: v3 [astro-ph.ga] 11 Jan 2019

Size: px
Start display at page:

Download "arxiv: v3 [astro-ph.ga] 11 Jan 2019"

Transcription

1 SUBMITTED TO APJ Preprint typeset using LATEX style emulateapj v. 01/23/15 COLD FILAMENTARY ACCRETION AND THE FORMATION OF METAL POOR GLOBULAR CLUSTERS AND HALO STARS NIR MANDELKER 1,2,3, PIETER G. VAN DOKKUM 2, JEAN P. BRODIE 4, FRANK C. VAN DEN BOSCH 2, DANIEL CEVERINO 5 Submitted to ApJ arxiv: v3 [astro-ph.ga] 11 Jan 2019 ABSTRACT We propose that cold filamentary accretion in massive galaxies at high redshift can lead to the formation of star-forming clumps in the halos of these galaxies without the presence of dark matter sub-structure. In certain cases, these clumps can be the birth places of metal poor globular-clusters (MP GCs). Using cosmological simulations, we show that narrow streams of dense gas feeding massive galaxies from the cosmic web can fragment, producing star-forming clumps. We then derive an analytical model for the properties of streams as a function of halo mass and redshift, and assess when these are gravitationally unstable, when this can lead to collapse and star-formation in the halo, and when it may result in the formation of MP GCs. For stream metalicities > 0.01Z, this is likely to occur at z>4.5. At z 6, the collapsing clouds have masses of M and the average stream pressure is 10 6 cm 3 K. The conditions for GC formation are met in the extremely turbulent eyewall at 0.3R v, where counter-rotating streams can collide, driving very large densities. Our scenario can account for the observed kinematics and spatial distribution of MP GCs, the correlation between their mass and metalicity, and the mass ratio between the GC system and the host halo. For MW mass halos, we infer that 30% of MP GCs could have formed in this way, the remainder likely accreted in mergers. Our predictions for GC formation along circumgalactic filaments at high-redshift are testable with JWST. Keywords: globular clusters: general galaxies: formation instabilities 1. INTRODUCTION The origin of globular clusters (GCs) has long challenged models of galaxy formation. GCs are bi-modal, with blue, metal-poor (MP), and red, metal-rich (MR), sub-populations (Larsen et al. 2001; Brodie & Strader 2006; Brodie et al. 2012). The distribution of metalicities for the two populations typically peak at [Fe/H] 1.5 and 0.5 respectively, though both peaks tend to shift to higher [Fe/H] with increasing galaxy luminosity (Brodie & Huchra 1991; Brodie & Strader 2006; Peng et al. 2006; Forte et al. 2009; Kruijssen 2014). In the Milky-Way, the populations are typically divided at [Fe/H] 1.1. MP GCs typically comprise of the total GC population, with the fraction even higher in low mass dwarfs (Strader et al. 2006; Peng et al. 2006). Both populations have comparable ages, to within measurement errors, roughly 12.5 Gyr ± 1 Gyr (e.g. Marín-Franch et al. 2009; VandenBerg et al. 2013; Forbes et al. 2015), though most models suggest that MP GCs formed on average earlier than MR ones (e.g. Forbes et al. 2015). This corresponds to GCs having formed before z 3, possibly even before the end of reionization. The mass functions of both MP and MR GCs are roughly the same. Both have a log-normal distribution, with typical masses in the range M and an average mass of (Brodie & Strader 2006; Wehner et al. 2008; Harris et al. 2013). The cutoff at the low mass end is thought to be due to disruption and evaporation of low mass GCs since their formation, while the initial mass function may have been a power law (Elmegreen & Efremov 1997; Elmegreen 2010; Baumgardt 1998; Fall & Zhang 2001; Kravtsov & Gnedin 2005; Prieto & Gnedin 2008; McLaughlin & Fall 2008; Kruijssen 2015). However, others have argued that the tidal forces experienced by GCs in the early Universe may have been much weaker than assumed, and thus that it is unlikely that a power law initial mass function could have been transformed into a log-normal distribution (Gieles & Renaud 2016; Renaud et al. 2017). Regardless, the surviving GCs should have undergone mass loss since their formation. Models attempting to explain chemical abundance anomalies in GCs by invoking multiple stellar populations require the mass at formation to be > 20 times more massive than present day masses (see Kruijssen 2014 for a recent review of such models). However, empirical models based on the observational consequences of such massive proto GCs suggest the mass loss ratio must be less than 10 (Boylan-Kolchin 2017), while analytical models of the physics of GC disruption predict mass loss ratios as low as 3 (Fall & Zhang 2001; Kruijssen 2015), which is also found in N-body simulations of cluster disruption (Webb & Leigh 2015). MP and MR GCs have comparable sizes, with half light radii of 2 3pc (Brodie & Strader 2006). 1 corresponding author: nir.mandelker@yale.edu 2 Department of Astronomy, Yale University, PO Box , New Haven, CT, USA 3 Heidelberger Institut für Theoretische Studien, Schloss- Wolfsbrunnenweg 35, Heidelberg, Germany 4 University of California Observatories, 1156 High Street, Santa Cruz, CA 95064, USA 5 Universität Heidelberg, Zentrum für Astronomie, Institut für Theoretische Astrophysik, Albert-Ueberle-Str. 2, Heidelberg, Germany The MP GCs are particularly enigmatic objects, with some very puzzling properties. Their spatial distribution often appears even more extended than the stellar halo (Strader et al. 2011; Forbes et al. 2012; Durrell et al. 2014; Kruijssen 2014), suggesting that most of them did not form inside their host galaxies. The fraction of MP GCs outside of 10kpc is much higher than in the central 5kpc of galaxies (Harris et al. 2006). By contrast, MR GCs fol-

2 2 GLOBULAR CLUSTERS AND COLD STREAMS low the galaxy light and are associated with the stellar disk and bulge, suggesting that they formed along with the thick disk and the bulge, perhaps during an unstable clumpy phase (Elmegreen et al. 2008; Shapiro et al. 2010; Kruijssen 2015; Renaud et al. 2017; Mandelker et al. 2017), or following wet mergers at intermediate redshifts (Ashman & Zepf 1992; Kravtsov & Gnedin 2005; Li & Gnedin 2014; Li et al. 2017). MP GCs exhibit a mass-metalicity relation, known as the blue tilt", such that more massive MP GCs are more metal rich (Strader et al. 2006; Peng et al. 2006; Spitler et al. 2006; Harris et al. 2006). This may be due to self-enrichment of massive MP GCs (Harris et al. 2006; Strader & Smith 2008; Bailin & Harris 2009), but the mechanism may depend on environment and does not appear universal (Spitler et al. 2006; Strader et al. 2006; Brodie & Strader 2006). No such relation is observed for MR GCs (Brodie & Strader 2006; Wehner et al. 2008). The kinematics of the two populations are also different, with the population of MP GCs showing more tangential orbits with significant apparent rotation, as opposed to the mostly radial orbits expected if they were mainly accreted (Pota et al. 2013, 2015a,b). The tangential anisotropy of MP GCs increases with distance from the halo center (Agnello et al. 2014). The MR GCs, on the other hand, have more mixed orbits (Pota et al. 2015b). One of the most puzzling properties of GCs is that the total mass of the GC system (GCS) in a galaxy is a near constant fraction of the dark matter halo mass, with a ratio of ± 0.28 dex (Hudson et al. 2014; Harris et al. 2015, 2017). This is in stark contrast to the highly nonlinear relation between a galaxy s stellar mass and halo mass (e.g. Yang et al. 2003; Behroozi et al. 2013). While a linear relation between GCS mass and halo mass exists for the total GC population, it appears mainly driven by the MP GCS (Harris et al. 2015). This relation holds over 5 orders of magnitude in galaxy mass, and in extreme environments, such as entire clusters of galaxies and Ultra-Diffuse Galaxies (UDGs) (Harris et al. 2017; van Dokkum et al. 2017). Some have suggested that this relation is a coincidence, resulting from a stellar mass dependent destruction efficiency for GCs combined with the non-linear stellar-to-halo mass relation (Kruijssen 2015, though see Fall & Chandar 2012 for evidence against a mass dependent destruction efficiency for clusters in the local Universe) or as a result of hierarchical galaxy assembly and the central limit theorem (El-Badry et al. 2018). However, many others have pointed out that this is suggestive of a link between GC formation and halo assembly at high redshift (e.g Spitler & Forbes 2009; Harris et al. 2017; Boylan-Kolchin 2017). Recent measurements indicate that the radial extent of GCSs, as measured by their half-number radii, is a constant fraction of the halo virial radius, R GC 0.04R v (Forbes 2017). This may be further evidence for an intimate connection betweent the properties of GCSs and their dark matter host halos. However, Hudson & Robison (2017) found a non-linear relation between the sizes of GCSs and the virial radii of their host halos, albeit with a smaller sample than Forbes (2017). Further observations are needed to clarify this point. Several classes of models have been envoked to account for the formation of MP GCs. Some models propose that they form at the centres of dark matter halos at the earliest stages of galaxy formation, prior to reionization (Peebles 1984; Rosenblatt et al. 1988; Moore et al. 2006). However, there is no dynamical evidence for dark matter in GCs (Moore 1996; Conroy et al. 2011). A second class of models predict that GCs formed within the gaseous halos surrounding massive galaxies in the early Universe, as opposed to in the halo centres, due to instabilities in the halo (e.g. Fall & Rees 1985; Cen 2001; Scannapieco et al. 2004). A third class of models suggests that MP GCs formed in dwarf galaxies in the early Universe, possibly as a result of major mergers (Kravtsov & Gnedin 2005; Li & Gnedin 2014; Li et al. 2017; Kim et al. 2017). These then merged onto larger galaxies and deposited their GCs in the halos of their new hosts (Ashman & Zepf 1992; Kravtsov & Gnedin 2005; Muratov & Gnedin 2010; Elmegreen et al. 2012; Kimm et al. 2016; Kim et al. 2017; Renaud et al. 2017). The similarity between properties of GCs and those of young massive clusters (YMCs) in the local Universe has led to the suggestion that GCs may be the descendants of ordinary YMC formation at high redshift (Elmegreen & Efremov 1997; Kravtsov & Gnedin 2005; Prieto & Gnedin 2008; Kruijssen 2014, 2015, though see also Renaud et al. 2017). Kruijssen (2015) argues that GC formation is a two stage process, beginning with a rapid-disruption phase in the high-pressure environments of high redshift discs until mergers cause them to migrate out into the halo, followed by slow evaporation in the halos. While this model is able to reproduce many observed properties of GCs and GCSs assuming that all GCs formed at z 3, it is not at all clear that all GCs formed inside galaxy discs, and other formation mechanisms should be explored (see discussion in Kruijssen 2014). In particular, the observed connection between GCSs and their dark-matter host halos warrents further investigation as to whether such a relation could have existed at their formation. Recently, an empirical model has been proposed where MP GCs form directly in their dark matter host halos at z > 6 in direct proportion to the host halo mass, and then undergo subsequent hierarchical merging of halos and of GCSs (Boylan-Kolchin 2017). It was shown that this can yield z = 0 GCS masses that are consistent with observations, though no physical mechanism was proposed for the formation of GCs in this way. In this paper, we explore a new formation channel for MP GCs directly in the halos of massive galaxies at z 6. This is similar in spirit to the second class of models described above, but motivated by our new understanding of gas accretion and the structure of the circumgalactic medium (CGM) in massive galaxies at high redshift. Such galaxies are predicted to be fed by narrow, dense streams of cold, metal poor gas ( 2). We propose that these streams can become gravitationally unstable, leading to the formation of massive star-forming clumps in the halos of such galaxies, and in certain cases to the formation of MP GCs. The remainder of this paper is organised as follows. In 2 we review some of the theoretical background and observational evidence for filamentary accretion at high redshift. In 3 we use a cosmological simulation to illustrate that streams can form bound, star-forming clumps not associated with a merging dark-matter halo, and discuss some of their properties. We then discuss recent observations that are suggestive of such stream fragmentation. In 4 we discuss stream fragmentation analytically. We begin by estimating the characteristic radii, densities, and turbulent Mach numbers of streams as a function of halo mass and redshift. We then explore whether the streams are gravitationally unstable, and whether they can cool and form stars before reaching the

3 MANDELKER ET AL. 3 central galaxy. Finally, we speculate when this may lead to GC formation. In 5 we summarize our model and present specific predictions regarding the properties of MP GCs and GC systems. We discuss our results and propose avenues for future work in 6. For the model presented in 4 as well as the halo mass histories shown in Fig. 1, we adopt cosmological parametersω m = 0.3,Ω Λ = 0.7, H 0 = 70kms 1 Mpc 1, and a Universal baryon fraction f b = FILAMENTARY ACCRETION - THEORY AND OBSERVATIONS In this section we review the theoretical background of, and the observational evidence for, the existence of cold streams around massive halos, their properties, and how these relate to the assumptions of our model. The most massive haloes at any epoch lie at the nodes of the cosmic web, and are penetrated by cosmic filaments of dark matter (e.g. Bond et al. 1996; Springel et al. 2005; Dekel et al. 2009a). We refer to such halos as stream fed halos. They represent high-sigma peaks, much more massive than the Press-Schechter mass, M ν=1, of typical haloes at that time (Press & Schechter 1974). A crude upper limit for the minimal mass of a stream fed halo is that of a 2 σ peak in the cosmic density field, M ν=2 (Dekel & Birnboim 2006; Dekel et al. 2009a). In Fig. 1, we show both M ν=1 and M ν=2 as a function of redshift, computed using Colossus (Diemer & Kravtsov 2015; Diemer 2015). We also show the average mass evolution of halos with different z = 0 masses, computed by integrating equation 2 from Fakhouri et al. (2010) for the mean mass accretion rate onto halos as a function of redshift. Lower mass halos drop below M ν=2 at higher redshift, and thus cease being stream fed at earlier times. For example, a M halo at z = 0 ceases to be stream fed at z 5, while a M halo remains stream fed until z 2. At z > 2 the cooling time in sheets and filaments is shorter than the Hubble time (Mo et al. 2005). Furthermore, in all but the most massive clusters and their progenitors, the gas flowing along these dark matter filaments is predicted to be unable to support a stable accretion shock at the filament edge (Dekel & Birnboim 2006; Birnboim et al. 2016). Intergalactic gas that accretes onto the filaments from sheets and voids remains dense and cold, T 10 4 K, as it free-falls towards the filament axis and settles in a narrow dense stream. Even in halos above M shock M, which contain hot gas at the virial temperature, the gas in streams is expected to remain cold and penetrate efficiently through the hot CGM onto the central galaxy (Dekel & Birnboim 2006). The above analytical picture is supported by cosmological simulations (Kereš et al. 2005; Ocvirk et al. 2008; Dekel et al. 2009a; Ceverino et al. 2010; Faucher-Giguère et al. 2011; van de Voort et al. 2011; Tillson et al. 2015; Nelson et al. 2016). In these simulations, cold streams with widths of a few to ten percent of the virial radius penetrate deep into the halo. Many global properties of the streams and their interaction with the CGM and with galaxies can be deduced from simulations and compared to observations, as detailed below. Although cold streams have not been detected directly, there is mounting circumstantial observational evidence for their existence. Cosmological simulations indicate that the streams maintain roughly constant inflow velocities as they travel from the outer halo to the central galaxy (Goerdt & Ceverino 2015). The constant velocity, as op- Figure 1. Redshift evolution of different halo masses. The black line shows the 2 σ mass, M ν=2, above which halos are likely to be stream fed. The grey line shows the Press-Schechter mass, M ν=1, for comparison. Colored lines show the average evolution for halos with z = 0 halo masses log(m v/m ) = 10, 11, 12, and 13. The lines are solid when M v > M ν=2 and the halos are stream fed, and become dashed when the virial mass drops below M ν=2. More massive halos remain stream fed until later times. posed to the expected gravitational acceleration, suggests energy loss into radiation which may be observed as Lyman-α cooling emission (Dijkstra & Loeb 2009; Goerdt et al. 2010; Faucher-Giguère et al. 2010). Radiative transfer models suggest that the total luminosity and the spatial structure of the emitted radiation appear similar to Lyman-α blobs observed at z > 2 (Steidel et al. 2000; Matsuda et al. 2006, 2011). Radiative transfer models also show that a central quasar can power the emission by supplying seed photons which scatter inelastically within the filaments, producing Lyman-α cooling emission that extends to several hundred kpc and appears similar to observed structures (Cantalupo et al. 2014). Recent observations using the MUSE integral-field instrument suggest that such extended Lyman-α emitting nebulae are ubiquitous around the brightest quasars at z 3.5 (Borisova et al. 2016; Vernet et al. 2017). The cold streams consist mostly of neutral Hydrogen and should also be visible in absorption. They can account for observed Lymanlimit systems (LLSs) and damped Lyman-α systems (DLAs) (Fumagalli et al. 2011; Goerdt et al. 2012; van de Voort et al. 2012). Observations using absorption features along quasar sight-lines to probe the CGM of massive SFGs at z 1 2 reveal low-metalicity, co-planar, co-rotating accreting material (Bouché et al. 2013, 2016), providing further observational support for the cold-stream paradigm. Strong Lyman-α absorption has also been detected in the CGM of massive submillimeter galaxies (SMGs) at z 2 (Fu et al. 2016). The streams are predicted to play a key role in the buildup of angular momentum in high-z disk galaxies (Pichon et al. 2011; Kimm et al. 2011; Stewart et al. 2011, 2013; Codis et al. 2012; Danovich et al. 2012, 2015). This is due both to vorticity within the filaments that spins up dark matter halos (Codis et al. 2012, 2015; Laigle et al. 2015),

4 4 GLOBULAR CLUSTERS AND COLD STREAMS Figure 2. Filamentary accretion in a cosmological zoom-in simulation from the VELA simulation suite. We show surface densities of dark matter (left), gas (centre), and stars younger than 100Myr (right). The top and bottom rows are 100kpc and 30kpc on a side, respectively. Each panel is oriented face on to the central disk, and the integration depth of all panels is 40kpc. The solid circles in the top row mark the halo virial radius, R v, while the dashed circles in the top and bottom rows mark 0.3R v. The galaxy is fed by three dense, narrow gas streams embedded within thicker dark matter filaments. These streams lie in a plane that extends beyond R v, and remain coherent outside of 0.3R v. In the inner halo, the streams can fragment into dense clumps not associated with any dark matter overdensity, where star-formation occurs. Two such clumps are circled in the bottom row. Such fragmentation can potentially lead to the formation of MP GCs far from the centres of dark matter halos. and also to an impact parameter of the streams with respect to the galaxy centre, typically 0.3R v (Kereš & Hernquist 2009; Danovich et al. 2015; Tillson et al. 2015). Simulations show streams that remain cold and coherent outside of 0.3R v, inside of which a messy interaction region is seen, with strong shocks and highly turbulent flow, where the streams collide, fragment, and experience strong torques before spiralling onto the disk in an extended ring-like structure (Ceverino et al. 2010; Danovich et al. 2015). Observations of the inner regions of massive, M v > M, halos at z 3 performed with the Cosmic Web Imager (CWI) have revealed extended gaseous structures with large angular momentum (Martin et al. 2014a,b). While only a handful of such cases have been observed thus far, their structure and kinematics appear very simiar to predictions from cosmological simulations of the kinematics of cold streams (Danovich et al. 2015). Similar kinematic features have been detected in absorption studies of the CGM of massive SFGs at z 1 2 (Bouché et al. 2013, 2016). In order to illustrate some of the concepts discussed above, we show in Fig. 2 a snapshot of a cosmological zoom-in simulation from the VELA simulation suite (Ceverino et al. 2014; Zolotov et al. 2015). The simulation was run with the adaptive-mesh refinement (AMR) code ART (Kravtsov et al. 1997; Kravtsov 2003; Ceverino & Klypin 2009). The code incorporates gas and metal cooling, UV-background photoionization with self-shielding in gas with Hydrogen number densities n > 0.1cm 3, stochastic star formation, gas recycling, stellar winds and metal enrichment, thermal feedback from supernovae (Ceverino et al. 2010) and feedback from radiation pressure (Ceverino et al. 2014). The grid is refined using a quasi-lagrangian strategy based on the total mass within a cell, up to a maximal resolution of pc (physical) at all times, though the resolution in the outer halo can be significantly lower. Details regarding the simulation method and its limitations can be found in Mandelker et al. (2017). In Fig. 2 we show galaxy V19 (see table 1 of Mandelker et al. 2017) at z = At this time, V19 has a virial mass of M, a virial radius of 26kpc, and a stellar mass within 0.1R v of M. The last significant merger was a 1 : 4 merger at z 8, roughly 300Myr before the snapshot shown.

5 MANDELKER ET AL Figure 3. Mass profiles for the gas (blue), stars (red), young stars (age< 100Myr, green) and dark matter (black) in the left-hand clump marked in Fig. 2. The profiles are truncated at the tidal radius of the clump, r tid 1.4kpc. Thin dot-dashed lines show the total mass profiles while solid lines show the bound mass. Note that all the bound stars are young stars, so the solid green line lies on top of the solid red line. No dark matter is bound to the clump, while > 70% of the gas and stars are bound. Within the half mass radius of r h 450pc, all of the gas and > 90% of the stars are bound. The mean Hydrogen number density in the inner 100pc is 7.1cm 3, while within r h it is 1.6cm 3. In Fig. 2, we show maps of the surface density of dark matter (left), gas (centre), and stars younger than 100 Myr (right). The panels in the top and bottom rows are 100kpc and 30 kpc across respectively. Concentric solid and dashed circles mark R v and 0.3R v respectively. The integration depth of all panels is 40kpc, and they are oriented perpendicular to the angular momentum of the central star-forming disk (see Mandelker et al for details on defining this plane). There are three prominent gas streams extending beyond the halo virial radius which seem to lie in a plane, the stream plane. While the existence of such a plane is a generic feature in cosmological simulations, it does not typically coincide with the plane defined by the disk angular momentum as it does in this case (Danovich et al. 2012). These gas streams lie at the centres of much wider dark matter filaments. The dark matter filaments are difficult to detect within the virial radius, while the gas streams remain prominent and coherent until reaching the interaction region at 0.3R v. 3. FRAGMENTATION AND STAR-FORMATION IN STREAMS 3.1. Stream Fragmentation in Cosmological Simulations Several previous studies have addressed the possibility of clumpy accretion along streams and its potential effects on galaxy formation and disc instability (e.g Dekel et al. 2009a,b; Genel et al. 2012; Goerdt et al. 2015). Other studies addressed the possibility of clumps forming due to thermal instabilities in massive halos or in streams and its potential effect on heating the CGM in proto-clusters (Dekel & Birnboim 2008; Birnboim & Dekel 2011). However, the formation of baryonic clumps within streams that are not associated with dark matter halos has not been studied in cosmological simulations (though Pallottini et al do show one such example). In Fig. 2, there are several dense clumps in the gas streams in the inner 0.5R v that do not appear to be associated with dark matter sub-halos. Two such clumps are highlighted with circles in the bottom panels. These clumps are forming stars despite not being associated with any dark matter overdensities. In order to verify to what extent the highlighted clumps are devoid of dark matter, we show in Fig. 3 the cumulative mass profiles of gas, stars, young stars (age < 100Myr), and dark matter of the left-hand clump (hereafter clump 1) in Fig. 2. The corresponding profiles for the right-hand clump (hereafter clump 2) are qualitatively similar. The profiles are centered on the peak of gas density and extend out to the tidal radius of the clump. The tidal radius of a clump falling on a radial trajectory into a massive halo is given by the implicit formula (Tormen et al. 1998) r tid = R c ( mc (r tid ) M h (R c ) 1 2 dlogm h (R c )/dlogr ) 1/3, (1) where m c (r) is the total mass of the clump interior to the clumpcentric radius r, R c is the distance of the clump from the center of the host halo, and M h (R) is the total mass of the host halo interior to the radius R. The deprojected distance of clump 1 from the halo center is R c 13kpc 0.5R v, and its tidal radius is r tid 1.4kpc. The total mass profiles interior to r tid are shown in Fig. 3 as thin dot-dashed lines. We then estimate for each gas cell, stellar and dark matter particle within r tid whether or not it is bound to the clump, by comparing its velocity to the escape velocity from the cell/particle position to the tidal radius, V esc (r) = ( rtid 2 r ) 1/2 Gm c (r) r 2 dr. (2) The profiles of bound mass are shown in Fig. 3 as solid lines. There is no dark matter bound to the clump. The total mass profile of dark matter roughly scales as m dm (r) r 3, indicating a constant background of dark matter from the host halo. On the other hand, over 70% of the gas and stellar mass interior to r tid is bound. The total bound baryonic mass of the clump is M and its half mass radius is 450pc. Within this radius, all of the gas and over 90% of the stars are bound, and the mean density is 2cm 3. The mass-weighted mean stellar age of the clump is 19Myr, yielding an average SFR of > 0.2M yr 1. In Fig. 4 we show the mass-weighted metalicity of the gas phase in the same projection as the bottom panels of Fig. 2. Within the cold streams the typical metalicities are [Z/H] 1.5, consistent with observed metalicities of MP GCs. Similarly low metalicities have been found in the simulations analyzed by Fumagalli et al. (2011); van de Voort & Schaye (2012); Ceverino et al. (2016). The mass-weighted mean stellar metalicity within clump 1 is 0.02Z. Clump 2 is qualitatively similar to clump 1. It has a tidal radius of 850pc, within which all the dark matter is unbound, while > 80% of the gas and stars are bound. Its gas mass is M, similar to clump 1, but its stellar mass is only M. Its mean stellar age is 10Myr, younger than clump 1, with a mean stellar metalicitiy of 0.025Z.

6 6 GLOBULAR CLUSTERS AND COLD STREAMS We stress that we are not claiming that these specific clumps represent MP GCs, which we have no hope of resolving in these simulations. In fact these clumps disrupt as they approach the central galaxy, depositing some of their stars in the halo and some of them in the disc. However, these examples highlight the fact that gas streams may become unstable and form stars and stellar clusters far from the centre of any dark matter halo. Beyond the origin of MP GCs, this can have implications for the build-up of stellar halos around massive galaxies, and for the frequency of Lyα emitters around massive galaxies at high redshift (Farina et al. 2017) Observational Evidence There is preliminary observational evidence for the fragmentation of cold streams. Radiative transfer models that attempt to reproduce the emission spectra of extended Lyman-α nebulae around luminous quasars at z 3 4 in halos of mass M v M, indicate that a large amount of gas is distributed in dense, compact clumps out to distances of at least 50 kpc (Cantalupo et al. 2014; Arrigoni Battaia et al. 2015; Borisova et al. 2016). Assuming that the emission is powered mainly by photoionization from the central quasar, the models place limits on the clump properties, suggesting densities of n > 3cm 3, sizes R < 20pc, temperatures T 10 4 K, metalicities Z < 0.1Z, and typical velocities of V > 500kms 1. While such clumps would be unresolved in cosmological simulations, we note that the mean density found in clump 1 (Fig. 3) is of the same order. Under these conditions, hydrodynamic instabilities induced by the motion of these clouds through the hot halo gas should disrupt them on very short timescales (Arrigoni Battaia et al. 2015). This tension may be alleviated if the clumps are not travelling within a hydrostatic hot halo, but rather within an inflowing cold stream. It has been argued that such clumps may originate from Kelvin Helmholtz Instabilities (KHI) (Mandelker et al. 2016; Padnos et al. 2018), Rayleigh Taylor Instabilities (RTI) (Kereš & Hernquist 2009), or thermal instabilities (Cornuault et al. 2016). Alternatively, they may be due to turbulent fragmentation in the streams ( 4.2). Several recent observations highlight the possibility of starformation occuring inside cold streams. Recent MUSE observations of a z 6.6 QSO, hosting a 10 9 M black-hole with an Eddington ratio of 0.7, have revealed the presence of a Lyα nebula out to distances of 10kpc from the QSO (Farina et al. 2017). This is 25% of the virial radius of a M halo at z = 6.6 (eq. 7). The inferred hydrogen volume density in the nebula is n H 6cm 3, similar to the densities in streams near 0.25R v at z 6.6 predicted by our model (eq. 24). Furthermore, the authors detected a Lyα emitter (LAE) at a projected distance of 12.5 kpc from the QSO, with a line-of-sight velocity difference of 560kms 1, comparable to the virial velocity of a M halo at z = 6.6 (eq. 8), located along the direction of the extended nebula. The inferred SFR in this LAE is 1.3M yr 1. Based on the QSO-galaxy cross-correlation function, the authors estimate that the probability of finding such a close LAE is < 10%. However, such a structure is a natural outcome of our model of star-forming clumps in gravitationally unstable streams, as detailed below. A similar example has been observed in an extended LAE near a bright QSO at z 3 (Rauch et al. 2013), where it was speculated that feedback from the central QSO triggered a burst of star formation at an inferred rate of 7M yr 1 in a Figure 4. Metalicity distribution in the gas streams and central halo. We show the mass-weighted average gas metalicity along the line of sight, with the same projection and integration depth as in the bottom row of Fig. 2. The gas streams feeding the central galaxy have typical metalicity values of Z. The clumps forming within the streams have similar metalicity values. The high metalicity material outside the streams is low density outflowing gas driven by stellar feedback in the central galaxy. nearby dwarf galaxy roughly 17kpc away from the QSO. We posit that such bursts of SF may occur directly in the dense, cold gas accreting onto galaxies without the need for a satellite dark matter halo or the trigger from a QSO. Finally, ALMA observations of a very massive, M M, galaxy at z 3.5 reveal a large structure of molecular gas extending out to 40kpc from the galaxy center (Ginolfi et al. 2017). This structure has a gas mass of , about 60% of which does not appear to be associated with either the central galaxy or its satellites. Such a large molecular gas mass cannot be accounted for by tidal stripping of satellite galaxies, and kinematic analysis does not reveal coherent rotation in the extended structure. This extended structure is also detected in continuum thermal emission, with an inferred SFR of 0.1M yr 1. The authors further detect gas rich systems on scales of up to 500kpc oriented along the same direction as the extended molecular source, with gas masses in the range M and SFRs of M yr 1. The authors interpret these results as gravitational collapse and fragmentation leading to starformation in the dense inner part of a cold stream feeding the central galaxy. A similar finding was reported in the Spiderweb galaxy, a massive proto-cluster at z 2.2 observed with the Australia Telescope Compact Array (ATCA) and the VLA (Emonts et al. 2016). These authors detected M of star-forming molecular gas within an extended Ly-α halo up to distances of 100 kpc from the central galaxy.

7 MANDELKER ET AL ANALYTIC ESTIMATES OF STREAM FRAGMENTATION Motivated by the results of the previous section, we present in this section an analytical study of gravitational instability and fragmentation in cold streams. In 4.1 we estimate the characteristic sizes and densities of cold streams. In 4.2 we identify several sources of turbulence in the streams and estimate the resulting turbulent Mach numbers. In 4.3 we assess the gravitational stability of streams and their characteristic fragmentation scales. In 4.4 we discuss cooling below 10 4 K and estimate when star-formation can occur in the collapsed clouds. Finally, in 4.5 we combine all the previous aspects of the model to assess when it may be possible to form MP GCs in streams. In the next section, 5, we summarize the main aspects of our model and highlight specific predictions for the properties of individual MP GCs and GC systems. A schematic illustration of our model, Fig. 7, and a table summarizing the main model parameters, Table 1, is presented here. Throughout, we normalize our results to streams feeding a M halo at z = 6, which corresponds to the main progenitor of a M halo at z = 0, i.e. a Milky-Way progenitor (Fig. 1). However, our model applies to all halos with M v > M ν=2, and we provide scalings with halo mass and redshift for all derived quantities Characteristic Densities and Sizes of Cold Streams We assume that the mass flux along the streams, which have a cylindrical or conical shape, is constant along their length until they reach the central halo (e.g. Dekel et al. 2009a). We thus have Ṁ s =πr 2 sρ s V s = M L,s V s. (3) where Ṁ s is the mass flux along the stream, r s is the stream radius,ρ s is its density, V s its velocity, and M L,s =πr 2 s ρ s is the mass per unit length of the stream, hereafter the line-mass. As detailed below, we allow the stream density and radius to vary with radial position within the halo. At z > 2, the mean accretion rate of total mass (baryons and dark matter) onto the virial radius of a halo with mass M v at redshift z is well fit by 6 (Fakhouri et al. 2010) Ṁ v 23M yr 1 M (1 + z)2.5 7, (4) where M 10 = M v /10 10 M and (1 + z) 7 = (1 + z)/7. The baryonic accretion rate onto the virial radius is given by multiplying eq. (4) by the universal baryon fraction, f b. At the high redshifts we are discussing we may assume that the accreted baryons are all gas. Cosmological simulations indicate that up to 50% of the accretion onto the halo is carried by one dominant stream, with typically two less prominent filaments carrying up to 10 20% each (Danovich et al. 2012). While these simulations focussed on M v M halos at z 2, we assume that similar fractions apply in stream fed halos at all redshifts. We adopt f s = 1/3 as our fiducial value for the fraction of the total accretion carried by a typical stream, and obtain for the gas accretion rate along such a stream Ṁ s f s f b Ṁ v 1.3M yr 1 f s,3 M (1 + z)2.5 7, (5) 6 While eq. (4) may slightly underpredict the accretion rate at z > 5 (van den Bosch et al. 2014), we note that a nearly identical formula can be derived analytically in the EdS regime at z > 1 (Dekel et al. 2013). where f s,3 = f s /(1/3) and we have used f b = Fluctuations in the total accretion rate, namely in the normalization of eq. (4), can be absorbed into f s yielding a plausible range of f s, (Dekel et al. 2013). The flow velocity along the stream may be written as V s =M v V v, (6) where V v = (GM v /R v ) 1/2 is the virial velocity of the halo, and M v is an effective Mach number", defined as the ratio of the stream velocity to the halo virial velocity. Cosmological simulations indicate that the streams maintain a roughly constant inflow velocity slightly below the virial value as they travel through the halo (Dekel et al. 2009a; Goerdt & Ceverino 2015). We assume M v 1, with an uncertainty of a factor of 2. The virial radius 7 and velocity of dark matter halos at z > 1 are given by (e.g. Dekel et al. 2013): R v 10kpc M 1/3 10 (1 + z) 1 7. (7) and V v 66kms 1 M 1/3 10 (1 + z)0.5 7, (8) Eq. (3) can be combined with eqs. (5)-(8) to yield the typical line-mass of streams feeding a halo with virial mass M v at redshift z: M L,s M kpc 1 M (1 + z) 2 7 f s,3 M v 1. (9) The line-mass of the host dark matter filament is given by dividing eq. (9) by the Universal baryon fraction, M L,Fil M kpc 1 M (1 + z)2 7 f s,3m v 1. (10) The characteristic radius of dark matter filaments as a function of their line-mass and redshift may be estimated by considering the expansion, turnaround and virialization of cylindrical top-hat perturbations in an expanding matter dominated Universe. This is analogous to the spherical collapse model for dark matter halos. Using this model, Fillmore & Goldreich (1984) derived the trajectories of collapsing cylindrical shells. They found an overdensity of 3.5 within the cylinder at turnaround (in the spherical collapse model, the overdensity at turnaround is 5.55), and that complete radial collapse of the shell occurs at roughly 2.5 times the turnaround time (as opposed to twice the turnaround time in the spherical collapse model). However, Fillmore & Goldreich (1984) did not discuss virialization of the filament. The virial theorem per unit length for an infinite cylinder is 2T = GML 2, (11) where T is the kinetic energy per unit length and M L is the mass per unit length (Chandrasekhar & Fermi 1953; Ostriker 1964). Together with the expression for the gravitational potential at radius r outside a cylinder with line-mass M L, Φ(r) = 2GM L ln(r/r 0 ), (12) where r 0 is a reference radius, one obtains a ratio of exp( 0.25) 0.78 between the virial radius and the turnaround radius in the cylindrical collapse model (this ratio is 0.5 in the spherical collapse model). This yields a virial 7 We define the halo virial radius as the radius of a sphere with mean density 18π times the mean Universal matter density, valid at z>1.

8 8 GLOBULAR CLUSTERS AND COLD STREAMS overdensity inside the collapsed cylinder of The mean density in a virialized filament is thus ρ v,fil 40Ω m ρ crit (1 + z) gr cm 3 (1 + z) 3 7, (13) where ρ crit gr cm 3 is the critical density of the Universe at z = 0. Combining this with eq. (10), yields the virial radius of a dark matter filament, r v,fil 8kpc M (1 + z) f 0.5 s,3 M v 0.5. (14) The resulting ratio of filament radius to halo virial radius is r v,fil /R v 0.8 M (1 + z) f 0.5 s,3 M v 0.5. (15) The relative width of filaments becomes smaller at later times, and has a very weak dependence on halo mass. For a M halo at z 6, this is roughly consistent with the most prominent filament seen on the left hand side of Fig. 2. For a M halo at z = 2, the typical ratio is 0.66, which appears consistent with high-resolution cosmological zoom-in simulations (e.g. Danovich et al. 2015; Nelson et al. 2016). The gas streams are significantly narrower than the dark matter filaments, since efficient cooling allows the gas to collapse towards the filament axis (Dekel & Birnboim 2006; Birnboim et al. 2016). However, the final radius of the gas streams and the mechanism that supports them against gravity has not been studied in detail yet. We expect the collapsed stream to be supported by a combination of thermal and turbulent pressure (see below), and rotation as evidenced by vorticity in the filament (Codis et al. 2012, 2015; Laigle et al. 2015; Birnboim et al. 2016). Self-consistently modelling all these sources of support is beyond the scope of the current paper, and will require detailed simulations of stream evolution. In appendix A, we show that assuming no rotation in the streams, equivalent to assuming that the streams are built by purely radial accretion of gas onto the centers of dark matter filaments, yields results that are inconsistent with both cosmological simulations and observations. Here, we adopt the opposite extreme and assume that the streams are largely supported by rotation, as has been suggested in the literature (e.g. Birnboim et al. 2016). This allows us to gain a crude estimate of the plausible sizes of streams, which turns out to be much more consistent with simulations and observations. We assume that the stream radius, r s, scales with the radius of the dark matter filament via a contraction factor,λ s, λ s r s /r v,fil. (16) We further assume that the dark matter filament can be approximated as an isothermal cylinder truncated at r v,fil. The density profile of an isothermal cylinder is (Ostriker 1964): ρ(r) =ρ c [1 + ( r/r h ) 2 ] 2, (17) whereρ c is the central density of the filament, and r h is its half mass radius. The filament line-mass profile is Inserting eq. (17) yields M L (r) = 2π r 0 ρ(r) r dr. (18) M L (r) = M L,Fil [1 ( 1 + (r/r h ) 2) 1 ], (19) 8 Harford & Hamilton (2011) quote a virial overdensity of 80, which is what one would infer if one assumes that the ratio of virial radius to turnaround radius for collapsed cylinders is 0.5, as it is for spheres. where M L,Fil =πr 2 h ρ c is the total line-mass of the filament. The associated circular velocity profile is V circ (r) = V ( r/rh ) [ 1 + ( r/r h ) 2 ] 1/2, (20) with V = (2GM L ) 1/2 the circular velocity at r. This can be related to the filament virial velocity using eq. (11), yielding V = 2V v,fil. If we assume that the specific angular momentum of the gas, j = r s V circ (r s ), is conserved during the gas contraction and that it is similar to that of the dark matter in the virialized filament, then λ s = 2 λ Fil V circ (r v,fil )/V circ (r s ), (21) where λ Fil = j/( 2r v,fil V v,fil ) is the filament spin parameter, defined here analogously to the halo spin parameter of Bullock et al. (2001). We denote 9 r v,fil /r h = β, and thus r s /r h =βλ s. Ifβλ s << 1 eq. (21) reduces to λ s [ 2/(1 +β 2 ) ] 1/4 λ1/2 Fil. (22) This yields λ s / λ 1/2 Fil 1, 0.8 or 0.37 for β = 1, 2, or 10 respectively. Eq. (22) relates the stream contraction factor to the filament spin parameter. It has been known for some time that the spin parameter of dark matter halos has a constant average value of λ H 0.035, independent of mass or time (Bullock et al. 2001). We assume that the filament spin parameter is likewise independent of mass or time, but with a smaller average value since the angular momentum of dark matter halos is predicted to originate from the combined angular momentum of the filament spin and orbit (impact parameter with respect to the halo centre) (Stewart et al. 2013; Danovich et al. 2015; Laigle et al. 2015). For λ Fil and β 1 10, we have λ s We hereafter assume as our fiducial value λ s 0.1, but caution that this is highly uncertain and could easily vary by a factor of a few. Cosmological simulations indicate that the streams assume a conical shape within the halo due to the gravitational attraction towards the halo center (e.g. Dekel et al. 2009a; van de Voort & Schaye 2012). We thus assume r s R within the halo, where R is the halocentric radius. Together with eqs. (14) and (16), this yields an expression for the stream radius, r s (R) 0.8kpc x v M (1 + z) Ł1 fs,3 0.5 M v 0.5, (23) where Ł1 = λ s /0.1 and x v = min(r/r v, 1). This is comparable, within a factor 2, to the size of the large stream seen in Fig. 2. We note that in shock-heated halos with M > v streams with radii r < s 0.05R v are not expected to reach the central halo, since they will be shredded by KHI (Mandelker et al. 2016; Padnos et al. 2018). This yields a practical lower limit of Ł1 > 0.5 in such massive halos. Eq. (23) can be used together with eq. (9) to obtain the average Hydrogen number density in the streams 10, n H,s 0.3cm 3 x 2 v (1 + z)3 7 Ł1 2, (24) which is independent of halo mass. At x v 0.5, the position of clump 1 in Fig. 2, this yields a density of 1.2cm 3, 9 β 1 is analogous to the concentration of spherical dark-matter halos. 10 For a primordial composition of Hydrogen and Helium, the corresponding mass density in the streams is roughly ρ s gr n H,s.

9 MANDELKER ET AL. 9 comparable to the density within the half mass radius of the clump, where all the gas is still bound (Fig. 3). When scaled to z 2, eq. (24) yields densities that are comparable to gas densities found in streams in cosmological simulations, which are in the range cm 3 (Goerdt et al. 2010; Faucher-Giguère et al. 2010; van de Voort & Schaye 2012). Combining eqs. (23) and (24) we can estimate the column density in a typical stream, cm 2 x 1 N H,s = 2r s n H,s v M (1 + z)2.5 7 Ł1 1 f 0.5 s,3 M v 0.5. (25) For a M halo at z 6, this corresponds to a surface mass density of 35M pc 2, within a factor of 2 of the typical gas surface densities in the streams in Fig. 2. Given their high column densities, the streams should be mostly self-shielded against the mean UV background radiation at all times. This has been found in simulations at z 2 (Goerdt et al. 2010; Faucher-Giguère et al. 2010) Turbulence in Streams In this section we try to estimate the turbulent velocities and turbulent Mach numbers in the streams. The temperature of the stream gas is T s 10 4 K, near the Lyman cooling floor, (Dekel & Birnboim 2006; Birnboim et al. 2016). The isothermal sound speed is thus c s = ( kb T s µ ) 1/2 8.2kms 1 T 0.5 4, (26) where T 4 = T s /10 4 K, k B is Boltzmann s constant, and µ 1.2m p is the mean molecular weight, with m p the proton mass. The chosen value for µ is valid for a nearly primordial composition of neutral gas. T 4 can vary in the range (Goerdt et al. 2010), absorbing any possible variation in µ as well. The first source of turbulence we consider is accretion of gas onto the streams from the large-scale pancakes within which they are embedded (Zel dovich 1970; Danovich et al. 2012), driven by the gravity of the streams and their host dark matter filaments. It is well established that such accretion generates turbulence (Klessen & Hennebelle 2010; Heitsch 2013; Clarke et al. 2017; Heigl et al. 2017). Based on the models of Klessen & Hennebelle (2010), Heitsch (2013) predicted the level of accretion driven turbulence in cylindrical filaments to be ( ) 1/3 σ acc = 2ǫr s V 2 Ṁ L,s, (27) acc M L,s where ǫ is an efficiency parameter which is inversely proprtional to the density contrast between the filament and the accreting material (Klessen & Hennebelle 2010), V acc is the accretion velocity onto r s, and Ṁ L,s is the mass accretion rate onto the stream. We estimate the expected radial accretion velocities onto the cold streams by considering the free-fall velocity of a cylindrical gas shell starting from rest at the cylindrical turnaround radius, 1.3r v,fil ( 4.1). Note that even if the gas has some net rotation velocity, as argued in 4.1, we are here only interested in the radial component. At r>r v,fil the gas is accelerated due to a constant line-mass M L,Fil, which is the total line-mass of the filament. The radial velocity at r v,fil is v r (r v,fil ) = 2 [ GM L,Fil ln(1.3) ] 1/2 (GML,Fil ) 1/2 = V v,fil, (28) where we have used eq. (12) for the gravitational potential outside a filament with line-mass M L,Fil. V v,fil = (GM L,Fil ) 1/2 is the virial velocity of the dark matter filament ( 4.1). At r v,fil > r> r s, the line-mass as a function of radius is given by eq. (19). This can be used to compute the change in potential from r v,fil to r s and thus the radial velocity at r s : ( ) 1 +β v 2 r (r s ) = v 2 2 r (r v,fil ) + 2GM L,Fil ln 1 + (βλ s ) 2, (29) where we recall that β = r v,fil /r h and λ s = r s /r v,fil. For our fiducial value of λ s = 0.1 and β = 1 10, we obtain v r (r s ) 1.5 3(GM L,Fil ) 1/2 2.3(GM L,Fil ) 1/2. Inserting eq. (10) into eq. (29) yields V acc 51kms 1 M (1 + z) 7 f 0.5 s,3 M v 0.5. (30) It is striking that at z 6 this is comparable to the virial velocity of the dark matter host halo, though it declines more rapidly with redshift. We note that the actual infall velocity onto the streams may be smaller that the free-fall velocity computed above if there is some dissipation mechanism acting on the gas as it flows towards the filament axis. Since V acc in eq. (30) is a factor 2 larger than the virial velocity of the dark matter filament, this uncertainty should be within a factor of < 2 and can be absorbed into the efficiency parameter ǫ, discussed below. The accretion rate onto the stream can be obtained by taking the time derivative of eq. (9) and inserting eq. (4) and the relation (1 + z) 7 (t/0.95 Gyr) 2/3. The result is Ṁ L,s 0.05M kpc 1 yr 1 M (1 + z)3.5 7 f s,3 M v [ M (1 + z) 7 ], (31) where the expression in square brackets is 1 for the halo masses and redshifts we are considering. Dividing eq. (31) by eq. (9) yields the specific accretion rate of gas onto the stream, Ṁ L,s M L,s 2.8Gyr 1 (1 + z) (32) Since Ṁ L,s 2πr s V acc ρ acc, where ρ acc is the density of the accreted gas, we may combine eqs. (23), (24), (30), and (31) to obtain the density contrast between the accreted gas and the stream gas: ρ acc ρ s 0.02 x 3 v Ł13. (33) Inside the halo, the stream gas is > 100 times denser than the accreted gas. This appears roughly consistent with a visual impression from cosmological simulations (e.g. Danovich et al. 2012, figure 7). Using numerical simulations and analytical arguments, Klessen & Hennebelle (2010) found that the efficiency of converting the inflow kinetic energy to turbulent kinetic energy, ǫ from eq. (27), is approximatelyǫ ρ acc /ρ s, with an uncertainty of a factor of 3. We adopt as our fiducial value a somewhat conservativeǫ=0.01. Inserting this along with eqs. (30) and (32) into eq. (27) yields

10 10 GLOBULAR CLUSTERS AND COLD STREAMS an estimate for the turbulent velocities driven by accretion σ turb,acc 5.2kms 1 M (1 + z) 7(ǫ 1 x v Ł1) 1/3 ( f s,3 /M v ) 1/2, (34) where ǫ 1 = ǫ/0.01. Dividing by the sound speed (eq. 26) yields the turbulent Mach number associated with accretion M turb,acc 0.7 M (1 + z) 7(ǫ 1 x v Ł1) 1/3 ( f s,3 /[M v T 4 ]) 1/2. (35) For our fiducial parameters at z = 6 in the outer halo (x v > 0.5), the resulting turbulent Mach numbers for halos of mass log(m v /M ) = 10, 11, and 12 are 0.7, 1.7, and 4 respectively. Only for the most massive halos is the turbulence highly supersonic. In the inner halo, near 0.3R v, these values decrease by a factor of At z 2 the turbulence is transonic for even the most massive halos. In addition to accretion, there are several other potential sources of turbulence in the streams. As mentioned in 2, in massive halos with M > v M that contain hot gas at the virial temperature, the streams are susceptible to KHI caused by their interaction with the halo gas. The KHI results in oblique shocks within the stream that drive turbulence with Mach numbers of order M turb,khi 1 (Padnos et al. 2018). In the inner halo, cold streams in hot halos may also be unstable to RTI provided they have an impact parameter with respect to the halo centre, placing them above the low density gas in the potential well (Kereš & Hernquist 2009). Finally, cold streams in hot halos may also be thermally unstable (Cornuault et al. 2016), which can drive highly supersonic turbulence. We note that even in less massive halos, these processes may become relevant near the halo centre where hot gas ejected from the central galaxy due to feedback may form a hot corona (Sokolowska et al. 2017). Another source of turbulence and instabilities in the streams is satellite galaxies flowing along the streams towards the central galaxy. Simulations suggest that up to 30% of the accretion into massive galaxies at z 3 is associated with such satellite galaxies in the form of both major and minor mergers (Dekel et al. 2009a). These galaxies can locally stir up the gas in the streams, due both to their gravitational influence on and relative velocity with respect to the stream gas, potentially inducing large turbulent motions. Furthermore, winds ejected from these satellite galaxies into the stream gas can cause shocks and stir up turbulence, and have a profound influence on the structure of streams (Faucher-Giguère et al. 2016). Altogether we estimate that instabilities and feedback can drive turbulence in the streams with Mach numbers of 1 to a few, preferentially in massive halos with M v > M. When summed in quadrature to turbulent velocities driven by accretion, the turbulent Mach numbers in streams near R v may reachm turb,tot 1, 2 and 4 5 in halos of mass 10 10, and M respectively at z 6. However, by z = 3 these values should decrease by 50% Gravitational Instability and Fragmentation Self-gravitating filaments are unstable to local perturbations at wavelengths larger than the 3d Jeans length, provided this is smaller than the radial scale of the filament, even in the presence of rotation (Freundlich et al. 2014). The Jeans length is given by L J = [πc 2 eff/(gρ s )] 1/2, (36) where c 2 eff = c2 s +σ2 turb /3 represents the combined radial support due to thermal and turbulent pressure. Based on the discussion in the previous section, we have c eff /c s 1.1, 1.5, and 2.5 for streams in halos of mass 10 10, and M respectively at z 6. At lower redshifts, the contribution of turbulent support decreases further. Therefore, in the following discussion we simply adopt c eff = c s, and comment where relevant on how this may underestimate the resulting fragmentation scales. This has the advantage of examining the degree to which thermal pressure alone can support the streams against fragmentation, since streams are often modeled in the literature simply as isothermal cylinders at T s 10 4 K (e.g. Dekel & Birnboim 2006; Harford & Hamilton 2011; Mandelker et al. 2016). Using eqs. (23), (24) and (26) we may compute the ratio of the thermal Jeans length to the filament diameter, L J 2r s 1.4 M (1 + z) 1 7 f 0.5 s,3 M v 0.5. (37) Note that this is independent of λ s and of position within the halo. For a M halo, a stream carrying f s = 0.5 of the total accretion has a thermal Jeans length equal to the stream width. For more massive halos L J < 2r s, which means that local density perturbations can trigger gravitational collapse within the streams. It can easily be verified that adopting the turbulent Jeans length does not change this conclusion. For streams where L J < 2r s, the characteristic mass of gravitationally unstable clumps is given by the Jeans mass, M J (4π/3)ρ s (L J /2) M x v (1+z) Ł1. (38) Note that the thermal Jeans mass is independent of halo mass, and scales linearly with halocentric radius. At z = 6, the turbulent Jeans mass is larger by a factor of 3 and 15 in streams feeding halos of mass and M. At lower redshifts the difference between the thermal and turbulent Jeans masses becomes smaller. In appendix B we show that the effects of the tidal fields induced by the host halo and the host darkmatter filament on the fragmentation scale are both negligible. In order to ascertain whether such a Jeans unstable cloud will have time to collapse before the stream reaches the central galaxy, we compare the free fall time, t ff = [3π/(32Gρ s )] 1/2, to the inflow time from a given radius, t inflow = R/V s. We find t ff t inflow 0.6 Ł1M v. (39) This is independent of halo mass, redshift, or position within the halo. Wherever the perturbation is seeded, it will collapse in roughly half the time it takes to reach the central galaxy. For long wavelength perturbations, larger than the filament diameter, the above local stability criterion cannot be used, and we must instead examine the global stability of the filament. A self-gravitating isothermal filament is unstable to global axisymmetric perturbations if its line-mass is larger than a critial value M L,crit = 2c 2 s/g (e.g. Ostriker 1964; Inutsuka & Miyama 1992). This can be thought of as a filamentary Jeans line-mass, above which thermal pressure cannot prevent global radial collapse of the filament. The ratio of the line-mass in a typical stream to this critical value is M L,s 0.6 M (1 + z) 2 7 f s,3 M 1 v T4 1. (40) M L,crit

11 MANDELKER ET AL. 11 At z = 6, streams feeding halos more massive than M are supercritical, and thus globally unstable. At z = 3, the critical halo mass for unstable streams is M, smaller than M ν=2 at that redshift (Fig. 1). The characteristic collapse time for a supercritical filament is t coll 3(4πGρ s ) 1/2 (Inutsuka & Miyama 1992; Heitsch 2013). Comparing this to the inflow time we find t coll t inflow 0.9 Ł1M v, (41) so the stream has time to globally fragment before reaching the central galaxy. This global instability results in the formation of dense cores separated by a few times the filament diameter (Inutsuka & Miyama 1992; Clarke et al. 2016). These cores then proceed to fragment on the local Jeans scale discussed above (Clarke et al. 2016, 2017). Our analysis thus suggests that most streams are supercritical with fragmentation times shorter than the halo crossing time. This is exacerbated further by the additional inwards gravitational force of the dark matter filament. This means that there must be some additional source of support in the streams which prevents catastrophic fragmentation in cosmological simulations. This could be rotation ( 4.1), turbulence ( 4.2), or artificial support caused by low resolution. However, Heitsch (2013) found that accretion-driven turbulence, which we expect to be the dominant source of accretion in most cases, can slow down the global collapse of supercritical filaments but it cannot halt it. The basic reason is that the line-mass of the stream grows faster than the resulting turbulent support, and therefore the stream remains super-critical. This has been confirmed by Clarke et al. (2017), who simulated a self-gravitating isothermal filament growing by accretion from its surroundings. They found that as long as the accretion flow itself was not highly turbulent, gravitational collapse and fragmentation occurs in two stages, first on large scales set by filamentary fragmentation, and then on small scales set by the local Jeans scale. To summarize, we find that most streams feeding massive halos should be gravitationally unstable to both short and long wavelength perturbations. The former result in direct three dimensional collapse on the Jeans scale, of clouds with masses given by eq. (38). The latter result first in two dimensional filamentary collapse, followed by subsequent three dimensional collapse on the Jeans scale, leading to collapsed clouds with the same mass as in the former case. Both of these processes can act in less than a virial crossing time, particularly if the stream is relatively narrow with Ł Cooling and Star Formation In order for star-formation to occur in the collapsing gas clumps, they must be able to cool from the initial stream temperatures of T s 10 4 K down to 10K. If the streams are indeed self shielded against the UV background then at 10 4 K they are mostly neutral, as assumed above. At metalicities of Z > 10 3 Z, and at the characteristic densities and temperatures of the streams, the dominant cooling process is emission in the [CII] 158µm line (Krumholz 2012; Pallottini et al. 2017). For a cloud of gas with a mean Hydrogen number density n H = n 0 cm 3, metalicity Z = 0.01Z Z 2, temperature T = 10 4 K T 4, and clumping factor C =< n 2 H > / n2 H = 10C 10, the ratio of the cooling time to the free-fall time of the cloud is (Krumholz 2012, equation 6) t cool t ff 9.1exp(0.009/T 4 ) Z 1 2 C 1 10 n 1/2 0 T 4. (42) For a collapsing cloud, this ratio decreases as the density increases during the collapse. However, since both the free-fall time and the cooling time are dominated by their initial stages near the onset of collapse, this initial ratio is representative of the final ratio. Inserting eq. (24) for the mean density in the streams and using T 4 1 this yields t cool t ff 16.8 Z 1 2 C 1 10 T 4 x v (1 + z) Ł1. (43) Note that this depends on position in the halo through the stream density. Multiplying eq. (43) by eq. (39) yields the ratio of cooling time in the streams to the inflow time, t cool 10.1 Z2 1 C10 1 T 4 x v (1 + z) Ł1 2 M v. (44) t inflow This ratio increases towards lower redshifts, as the streams become less dense. We define the cooling redshift, z cool as the redshift where the cooling time is equal to the inflow time, 1 + z cool 32.7 (Z 2 C 10 ) 2/3( T 4 M v Ł1 2) 2/3 x 2/3 v. (45) At z<z cool, the cooling time in the collapsing clump is longer than the inflow time to the halo center, and we do not expect much star-formation in the clump. However, at z>z cool, the clump may experience a burst of star-formation before reaching the central galaxy. Clumping factors of C 5 10 are not unreasonable at z 6 given the levels of turbulence and substructure in the streams. In fact observations suggest that the clumping factors may be even higher in very massive halos ( 3.2). We therefore assume C 10 > 0.5 at z 6, possibly approachingc 10 1 for M halos where the turbulent velocities are very large. However, as the turbulent velocities decrease towards lower redshift the associated clumping factors may decrease as well. For a relatively narrow stream with Ł1 = 0.5 near the inner halo at x v = 0.3, with a metalicity of Z 2 = 2 (Fig. 4) and a clumping factor of C 10 = 0.5, this yields z cool 4.8. At z = 6, this yields t cool 2.1t ff 0.6t inflow. In this case, the clump can cool and proceed to star-formation on a free-fall timescale, before reaching the central galaxy. Furthermore, since the initial cooling timescale is longer than the free-fall timescale of the cloud, we expect relatively large contraction factors for these clouds before the onset of star-formation. This can increase the mean densities in the clouds by considerable amounts compared to the mean densities in the streams given by eq. (24). It is also worth noting that such a cloud can cool and form stars before forming any appreciable fraction of H 2 (Krumholz 2012). In order to obtain z cool 2 in a stream with fiducial parameters, we require either larger metalicities closer to 0.1Z, or larger clumping factors. Overall, we predict star-formation to be less likely to occur in the streams outside the central galaxy at z < 4, unless the metalicities are Z, larger than indicated by most cosmological simulations (Fumagalli et al. 2011; van de Voort & Schaye 2012; Ceverino et al. 2016). In order for collapse to occur the clump must also dissipate its turbulent support. The initial clump radius is of order the stream radius (eq. 37). When accounting for turbulence in the

12 12 GLOBULAR CLUSTERS AND COLD STREAMS we assumed a relatively narrow stream, with Ł1 = 0.5. The turbulent Mach number is given by M turb,acc from eq. (35), summed in quadrature with an additional M 1 to account for instabilities and feedback (see 4.2). The white lines in Fig. 5 represent mass evolution tracks of halos with z = 0 masses log(m v /M ) = 10, 11, 12, 13, and 14. For the chosen parameters, the pressure in the streams exceeds 10 6 cm 3 K at z > 6 for all halo masses, and at z > 5 for the progenitors of the most massive halos, M v (z = 0) < Recall that the pressure plotted in Fig. 5, which represents the pressure in the streams prior to gravitational collapse, is not enough to maintain the streams in hydrostatic equilibrium when M L,s > M L,crit (eq. 40). The final pressure in the collapsed clouds prior to the onset of star-formation is thus likely to be even higher. We conclude that at z > 6 the pressure in the streams in the inner halo is large enough to enable formation of GCs. However, this may not be true outside of 0.3R v due to the strong dependence of the density and pressure on position within the halo, P ρ x 2 v Cluster Formation Efficiency Figure 5. Average pressure, P = ρc 2 eff, in the streams as a function of halo mass and redshift for parameters f s,3 =M v = T 4 = 1, Ł1 = 0.5, and x v = 0.3. This corresponds to our fiducial values for the accretion rate, velocity and temperature of the stream, but a relatively narrow stream in the inner halo at 0.3R v. The mean density is given by eq. (24), the sound speed by eq. (26), and the turbulent Mach number by eq. (35) with fiducial efficiency, ǫ 1 = 1, summed in quadrature with additional Mach 1 turbulence due to instabilities and feedback. Solid white lines mark the average mass histories of halos with present day masses log(m v(z = 0)/M ) = 10, 11, 12, 13, and 14, as marked. At z > 6, the pressure is P/k > B 10 6 cm 3 K, as required to form GCs. At lower redshifts the pressure falls below this threshold. Note that this represents the mean pressure in the stream prior to the onset of gravitational instability, while the pressure in the collapsed clouds is expected to be larger. Jeans length, this remains true even at M v M. The dissipation rate of turbulence over a length scale r s is equal to t diss r s /σ turb. Using eq. (34) as a proxy for the total turbulence in the clump, the ratio of this timescale to the inflow time can be evaluated: t diss 0.94 xv 1/3 Ł1 2/3 M v ǫ 1/3 1. (46) t inflow For a typical stream, the initial turbulence will dissipate in roughly the time it takes the stream to reach the central galaxy. At x v = 0.3 the ratio increases to 1.4 for Ł1 = 1, but if Ł1 = 0.5 it is Formation of MP GCs Gas Pressure in Streams We have established that streams at high redshift are unstable to fragmentation and that the cooling time is short enough to allow star-formation in the collapsing gas clouds. In order to form GCs in the collapsing clouds, they must reach high enough densities and pressures. GCs are expected to form only in very high pressure regions, with P/k B > 10 6 cm 3 K (Elmegreen & Efremov 1997; Kruijssen 2015). The combined thermal plus turbulent pressure in the streams prior to gravitational collapse is given by P ρ s c 2 eff (see 4.3). This is shown as a function of halo mass and redshift in Fig. 5. We assumed fiducial values for the stream velocity and temperature, the fraction of total accretion in the stream, and the efficiency of converting accretion energy into turbulent energy. In other words, M v = T 4 = f s,3 = ǫ 1 = 1. However, In addition to large pressures, the formation of a GC requires very high densities. A typical GC with a mass of and a half-mass radius of 3pc has a mean density within this radius of 880M pc cm 3. At z = 6, at 0.3R v, even a relatively narrow (and thus dense) stream with Ł1 = 0.5 has a typical density of 13cm 3 according to eq. (24), roughly a factor of 2000 below the densities in GCs. However, we cannot compare the mean density in the stream or in the pre-collapse cloud to the final GC density. Star-formation in a turbulent medium is a hierarchical process where the densest objects, namely bound stellar clusters and GCs, form at the highest density peaks within the cloud (e.g. Kruijssen et al. 2012; Hopkins 2013). Furthermore, there is mounting evidence in the local Universe that the densest stellar clusters are significantly denser than the densest gas clouds (Longmore et al. 2014; Walker et al. 2015), suggesting that massive clusters are formed via hierarchical merging of smaller stellar clusters embedded in the parent gas cloud. Within this framework, one can evaluate the fraction of star-formation occuring in bound clusters, referred to as the cluster formation efficiency, or CFE (e.g. Bastian 2008; Goddard et al. 2010), based on the density and pressure within the parent cloud 11 (Kruijssen 2012; Adamo et al. 2015; Johnson et al. 2016). For a M halo at z = 6 with the same parameters as in Fig. 5, the streams have pressure P/k B 10 6 cm 3 K, density ρ 0.44M pc 3, and column density Σ 100M pc 2. These properties are very similar to those of the Fornax model of Kruijssen (2015) (Table 1), which had P/k B = cm 3 K,ρ = 0.82M pc 3,Σ = 103M pc 2, and a CFE ofγ In our model, the density and pressure in the collapsed clouds prior to the onset of star-formation are likely to be much higher than the typical stream values, so CFE values of 0.4 or larger are entirely plausible. 11 There is still some debate in the literature regarding the origin of the CFE, with some claiming that it does not depend on the properties of the host galaxy or of the parent cloud (Fall & Chandar 2012; Chandar et al. 2015, 2017; Mulia et al. 2016). However, we here adopt the theoretical framework of Kruijssen (2012) whereby the CFE depends on the density and pressure in the parent cloud.

13 MANDELKER ET AL. 13 Figure 6. Mass fraction within collapsed clouds at densities ρ > ρ GC 880M pc 3, high enough to form GCs. As in Fig. 5, we assume stream parameters M v = f s,3 = T 4 =ǫ 1 = 1, and Ł1 = 0.5. The turbulence in the cloud is given by eq. (35) summed in quadrature with additional Mach 1 turbulence due to instabilities and feedback. The initial mean cloud density is given by eq. (24), but as the cloud contracts radially by a factor c = 5 (center panel) or 10 (left and right hand panels) its mean density increases by a factor c 3. The turbulence induces a density distribution given by eqs. (47) and (48) with b = 0.5. For the radial position in the halo, we show results for x v = 0.5 (left hand panel) and x v = 0.3 (center and right hand panels). Solid white lines mark the average mass histories of halos with present day masses log(m v(z = 0)/M ) = 10, 11, 12, 13, and 14, as marked. The vertical dashed lines mark z cool from eq. (45), the redshift above which the initial cooling time in the cloud prior to collapse is shorter than the inflow time of the stream to the central galaxy. For this equation we have assumed a clumping factor C 10 = 0.5 and a metalicity Z 2 = 2 (2% solar). The black dotted line in each panel marks the contour where the total mass of gas with ρ>ρ GC is M dense = M, assuming the total gas mass in the collapsed cloud is given by the turbulent Jeans mass, i.e. eq. (38) multiplied by a factor (1+M 2 turb /3)3/2. Above this curve the clouds have enough dense gas to form GCs. Due to the relatively low stream densities in the outer halo, x v > 0.5, the cooling time is long and we have z cool > 7. Furthermore, even with very large contraction factors of c = 10 and a narrow stream, there is not enough dense gas to form GCs except at very high redshifts z>7. With contraction factors of c = 5, the same is true even near the inner halo, at x v = 0.3. However, at x v = 0.3 with c = 10, roughly 10% of the cloud mass is dense enough to produce GCs in halos with virial masses down to 10 9 M at z>5.5, while z cool 4.8. Such large contraction factors likely require loss of angular momentum in the collapsing cloud, which can occur in the inner halo when counter-rotating streams collide (see text). The CFE can be used to determine the maximal cluster mass that forms within the parent cloud, M cl,max Γǫ SF M J whereǫ SF is the fraction of the cloud mass that turns into stars (Reina-Campos & Kruijssen 2017). For Γ 0.3, ǫ SF 0.1, and M J M (eq. 38), we infer a maximal cluster mass of order M which is again consistent with the maximal cluster mass in the Fornax model of Kruijssen (2015). For halos with M v > M at z 6, which are the progenitors of> M halos at z = 0, we expect larger CFE values due to stronger gravitational instability as well as larger Jeans masses due to stronger turbulence (see 4.3). Both of these will yield larger cluster masses Dense Gas Fraction Notwithstanding the above discussion, we defer a more detailed evaluation of the CFE and the maximal cluster mass in streams to future work, focussing here instead on the simpler question of under what conditions a collapsed cloud will have enough gas at high enough densities to lead directly to the formation of a GC. Based on the above discussion, this is a sufficient but not necessary condition for GC formation. As previously stated, a typical GC is 2000 times denser than a typical stream. For the mean density in the collapsed cloud to reach these values, a radial contraction factor of c = L J /R cl > 13 is required, where R cl is the radius of the collapsed cloud. At lower redshifts and at larger halocentric distances, the discrepancy is larger. While we have argued in the previous section that the relatively long cooling times in the clouds compared to their free-fall timescale can lead to significant contraction factors before the onset of star-formation, a contraction factor of c>10 may still be difficult to achieve due to angular momentum support. However, as described above it is not necessary for the mean density in the cloud to be as dense as a GC, but only that the highest density peaks are dense enough. The volumeweighted PDF of the mass overdensity in an isothermal turbulent medium is well described by a lognormal distribution (e.g. Federrath et al. 2010; Konstandin et al. 2012) dp/ds = 1 2πσ 2 ρ s exp where s = ρ/ ρ is the overdensity, and (( ) ln(s) + 0.5σ 2 2 ) ρ, (47) 2σ 2 ρ σ 2 ρ = ln(1 + b2 M 2 ), (48) wheremis the turbulent Mach number, and b depends on the ratio of compressive to solenoidal modes in the turbulence. For a natural" mixture of modes b 0.5, while it approaches 1 for more compressive forcing (Federrath et al. 2010). We can use this expression to estimate the fraction of mass that will be dense enough to form a GC, i.e. ρ > 880M pc 3. We assume that the cloud begins with mean density ρ s (eq. 24) and then contracts radially by a factor of c, so that its final density is ρ c 3 ρ s. The turbulence in the initial cloud is given by M turb,acc from eq. (35), summed in quadrature with an additional M 1 to account for instabilities and feedback, as in Fig. 5. As the cloud contracts, its internal turbulence may be amplified. In the absense of any dissipation or cooling, σturb 2 GM/R, so the turbulent Mach number scales as M turb c 1/2. However, in practice this maximal enhancement is rarely seen in simulations since turbulence dissipates more rapidly as the cloud contracts. As a conservative estimate, we ignore the possible enhancement of turbulence and

Violent Disk Instability at z=1-4 Outflows; Clump Evolution; Compact Spheroids

Violent Disk Instability at z=1-4 Outflows; Clump Evolution; Compact Spheroids Violent Disk Instability at z=1-4 Outflows; Clump Evolution; Compact Spheroids Avishai Dekel The Hebrew University of Jerusalem Santa Cruz, August 2013 stars 5 kpc Outline 1. Inflows and Outflows 2. Evolution

More information

The Magic Scale of Galaxy Formation: SNe & Hot CGM --> Compaction & BHs

The Magic Scale of Galaxy Formation: SNe & Hot CGM --> Compaction & BHs The Magic Scale of Galaxy Formation: SNe & Hot CGM --> Compaction & BHs Avishai Dekel The Hebrew University of Jerusalem & UCSC Silk 75, December 2017 A Characteristic Mass for Galaxy Formation Efficiency

More information

The formation and evolution of globular cluster systems. Joel Pfeffer, Nate Bastian (Liverpool, LJMU)

The formation and evolution of globular cluster systems. Joel Pfeffer, Nate Bastian (Liverpool, LJMU) The formation and evolution of globular cluster systems Joel Pfeffer, Nate Bastian (Liverpool, LJMU) Introduction to stellar clusters Open clusters: few - 10 4 M few Myr - few Gyr solar metallicity disk

More information

Outline. Walls, Filaments, Voids. Cosmic epochs. Jeans length I. Jeans length II. Cosmology AS7009, 2008 Lecture 10. λ =

Outline. Walls, Filaments, Voids. Cosmic epochs. Jeans length I. Jeans length II. Cosmology AS7009, 2008 Lecture 10. λ = Cosmology AS7009, 2008 Lecture 10 Outline Structure formation Jeans length, Jeans mass Structure formation with and without dark matter Cold versus hot dark matter Dissipation The matter power spectrum

More information

Galaxy Evolution & Black-Hole Growth (review)

Galaxy Evolution & Black-Hole Growth (review) Galaxy Evolution & Black-Hole Growth (review) Avishai Dekel The Hebrew University of Jerusalem & UCSC Delivered by Fangzhou Jiang Dali, China, November 2018 See also Claude-Andre s talk and Joel s talk

More information

Gas accretion in Galaxies

Gas accretion in Galaxies Massive Galaxies Over Cosmic Time 3, Tucson 11/2010 Gas accretion in Galaxies Dušan Kereš TAC, UC Berkeley Hubble Fellow Collaborators: Romeel Davé, Mark Fardal, C.-A. Faucher-Giguere, Lars Hernquist,

More information

Co-Evolution of Central Black Holes and Nuclear Star Clusters

Co-Evolution of Central Black Holes and Nuclear Star Clusters Co-Evolution of Central Black Holes and Nuclear Star Clusters Oleg Gnedin (University of Michigan) Globular clusters in the Galaxy median distance from the center is 5 kpc Resolved star cluster highest

More information

Feedback, AGN and galaxy formation. Debora Sijacki

Feedback, AGN and galaxy formation. Debora Sijacki Feedback, AGN and galaxy formation Debora Sijacki Formation of black hole seeds: the big picture Planck data, 2013 (new results 2015) Formation of black hole seeds: the big picture CMB black body spectrum

More information

Theoretical ideas About Galaxy Wide Star Formation! Star Formation Efficiency!

Theoretical ideas About Galaxy Wide Star Formation! Star Formation Efficiency! Theoretical ideas About Galaxy Wide Star Formation Theoretical predictions are that galaxy formation is most efficient near a mass of 10 12 M based on analyses of supernova feedback and gas cooling times

More information

Spin Acquisition, Violent Disks, Compaction and Quenching

Spin Acquisition, Violent Disks, Compaction and Quenching Spin Acquisition, Violent Disks, Compaction and Quenching Avishai Dekel The Hebrew University of Jerusalem July 2014 stars 5 kpc Three Provocative Questions concerning high-z massive galaxy formation 1.

More information

Origin of Bi-modality

Origin of Bi-modality Origin of Bi-modality and Downsizing Avishai Dekel HU Jerusalem Galaxies and Structures Through Cosmic Times Venice, March 2006 Summary Q: z

More information

Gaia Revue des Exigences préliminaires 1

Gaia Revue des Exigences préliminaires 1 Gaia Revue des Exigences préliminaires 1 Global top questions 1. Which stars form and have been formed where? - Star formation history of the inner disk - Location and number of spiral arms - Extent of

More information

Three comments on High-z Galaxy Formation. Avishai Dekel The Hebrew University of Jerusalem

Three comments on High-z Galaxy Formation. Avishai Dekel The Hebrew University of Jerusalem Three comments on High-z Galaxy Formation Avishai Dekel The Hebrew University of Jerusalem August 2014 Outline 1. Angular momentum: buildup in 4 phases 2. Violent disk instability: Nonlinear, Stimulated

More information

Nir Mandelker, H.U.J.I.

Nir Mandelker, H.U.J.I. Compressive vs Solenoidal Turbulence and Non-Linear VDI Nir Mandelker, H.U.J.I. IAU Symposium 319, August 11 2015 Collaborators: Avishai Dekel, Shigeki Inoue, Daniel Ceverino, Frederic Bournaud, Joel Primack

More information

Galaxy Formation and Evolution

Galaxy Formation and Evolution Galaxy Formation and Evolution Houjun Mo Department of Astronomy, University of Massachusetts 710 North Pleasant Str., Amherst, MA 01003-9305, USA Frank van den Bosch Department of Physics & Astronomy,

More information

Formation and cosmic evolution of supermassive black holes. Debora Sijacki

Formation and cosmic evolution of supermassive black holes. Debora Sijacki Formation and cosmic evolution of supermassive black holes Debora Sijacki Summer school: Black Holes at all scales Ioannina, Greece, Sept 16-19, 2013 Lecture 1: - formation of black hole seeds - low mass

More information

High-z Galaxy Evolution: VDI and (mostly minor) Mergers

High-z Galaxy Evolution: VDI and (mostly minor) Mergers High-z Galaxy Evolution: VDI and (mostly minor) Mergers Avishai Dekel The Hebrew University of Jerusalem UCSC, August 2012 Outline: in-situ (VDI) and ex-situ (mergers) 1. Cold streams: smooth and clumpy

More information

Feeding High-z Galaxies from the Cosmic Web

Feeding High-z Galaxies from the Cosmic Web Feeding High-z Galaxies from the Cosmic Web Avishai Dekel The Hebrew University of Jerusalem Jerusalem Winter School 2012/13 Lecture 1 z=0 z=2 z=8 Cosmological simulations Toy modeling Oxford Dictionary:

More information

Formation of z~6 Quasars from Hierarchical Galaxy Mergers

Formation of z~6 Quasars from Hierarchical Galaxy Mergers Formation of z~6 Quasars from Hierarchical Galaxy Mergers Yuexing Li et al Presentation by: William Gray Definitions and Jargon QUASAR stands for QUASI-stellAR radio source Extremely bright and active

More information

Recent Progress in Modeling of Galaxy Formation. Oleg Gnedin (University of Michigan)

Recent Progress in Modeling of Galaxy Formation. Oleg Gnedin (University of Michigan) Recent Progress in Modeling of Galaxy Formation Oleg Gnedin (University of Michigan) In current simulations, galaxies look like this: 10 kpc Disk galaxy at z=3: stars, molecular gas, atomic gas (Zemp,

More information

ASTR 610 Theory of Galaxy Formation Lecture 14: Heating & Cooling

ASTR 610 Theory of Galaxy Formation Lecture 14: Heating & Cooling ASTR 610 Theory of Galaxy Formation Lecture 14: Heating & Cooling Frank van den Bosch Yale University, spring 2017 Heating & Cooling In this lecture we address heating and cooling of gas inside dark matter

More information

Ionization Feedback in Massive Star Formation

Ionization Feedback in Massive Star Formation Ionization Feedback in Massive Star Formation Thomas Peters Institut für Theoretische Astrophysik Zentrum für Astronomie der Universität Heidelberg Ralf Klessen, Robi Banerjee (ITA, Heidelberg) Mordecai-Mark

More information

Mergers and Mass Assembly of Dark Matter Halos & Galaxies

Mergers and Mass Assembly of Dark Matter Halos & Galaxies Mergers and Mass Assembly of Dark Matter Halos & Galaxies Chung-Pei Ma Onsi Fakhouri James McBride (UC Berkeley) Mike Boylan-Kolchin (MPA --> Southern UC) Claude-Andre Faucher-Giguere Dusan Keres (Harvard

More information

Astro-2: History of the Universe

Astro-2: History of the Universe Astro-2: History of the Universe Lecture 13; May 30 2013 Previously on astro-2 Energy and mass are equivalent through Einstein s equation and can be converted into each other (pair production and annihilations)

More information

The Iguaçu Lectures. Nonlinear Structure Formation: The growth of galaxies and larger scale structures

The Iguaçu Lectures. Nonlinear Structure Formation: The growth of galaxies and larger scale structures April 2006 The Iguaçu Lectures Nonlinear Structure Formation: The growth of galaxies and larger scale structures Simon White Max Planck Institute for Astrophysics z = 0 Dark Matter ROT EVOL Cluster structure

More information

Stream-Driven Galaxy Formation at High Redshift

Stream-Driven Galaxy Formation at High Redshift Stream-Driven Galaxy Formation at High Redshift Avishai Dekel The Hebrew University of Jerusalem KooFest, Santa Cruz, August 2011 Outline 1. Streams in pancakes from the cosmic web (Hahn) 2. Is angular

More information

The First Galaxies. Erik Zackrisson. Department of Astronomy Stockholm University

The First Galaxies. Erik Zackrisson. Department of Astronomy Stockholm University The First Galaxies Erik Zackrisson Department of Astronomy Stockholm University Outline The first galaxies what, when, why? What s so special about them? Why are they important for cosmology? How can we

More information

Killing Dwarfs with Hot Pancakes. Frank C. van den Bosch (MPIA) with Houjun Mo, Xiaohu Yang & Neal Katz

Killing Dwarfs with Hot Pancakes. Frank C. van den Bosch (MPIA) with Houjun Mo, Xiaohu Yang & Neal Katz Killing Dwarfs with Hot Pancakes Frank C. van den Bosch (MPIA) with Houjun Mo, Xiaohu Yang & Neal Katz The Paradigm... SN feedback AGN feedback The halo mass function is much steeper than luminosity function

More information

Survey of Astrophysics A110

Survey of Astrophysics A110 Goals: Galaxies To determine the types and distributions of galaxies? How do we measure the mass of galaxies and what comprises this mass? How do we measure distances to galaxies and what does this tell

More information

Galaxies and Cosmology

Galaxies and Cosmology F. Combes P. Boisse A. Mazure A. Blanchard Galaxies and Cosmology Translated by M. Seymour With 192 Figures Springer Contents General Introduction 1 1 The Classification and Morphology of Galaxies 5 1.1

More information

Veilleux! see MBW ! 23! 24!

Veilleux! see MBW ! 23! 24! Veilleux! see MBW 10.4.3! 23! 24! MBW pg 488-491! 25! But simple closed-box model works well for bulge of Milky Way! Outflow and/or accretion is needed to explain!!!metallicity distribution of stars in

More information

Galaxies 626. Lecture 5

Galaxies 626. Lecture 5 Galaxies 626 Lecture 5 Galaxies 626 The epoch of reionization After Reionization After reionization, star formation was never the same: the first massive stars produce dust, which catalyzes H2 formation

More information

arxiv:astro-ph/ v1 4 Jan 1995

arxiv:astro-ph/ v1 4 Jan 1995 Cosmological Origin of Quasars ABRAHAM LOEB Astronomy Department, Harvard University 60 Garden St., Cambridge, MA 02138 Contribution to the Texas Symposium, Munich, Dec. 1994 arxiv:astro-ph/9501009v1 4

More information

8.1 Structure Formation: Introduction and the Growth of Density Perturbations

8.1 Structure Formation: Introduction and the Growth of Density Perturbations 8.1 Structure Formation: Introduction and the Growth of Density Perturbations 1 Structure Formation and Evolution From this (Δρ/ρ ~ 10-6 ) to this (Δρ/ρ ~ 10 +2 ) to this (Δρ/ρ ~ 10 +6 ) 2 Origin of Structure

More information

Origin and Evolution of Disk Galaxy Scaling Relations

Origin and Evolution of Disk Galaxy Scaling Relations Origin and Evolution of Disk Galaxy Scaling Relations Aaron A. Dutton (CITA National Fellow, University of Victoria) Collaborators: Frank C. van den Bosch (Utah), Avishai Dekel (HU Jerusalem), + DEEP2

More information

Angular Momentum Acquisition in Galaxy Halos

Angular Momentum Acquisition in Galaxy Halos Angular Momentum Acquisition in Galaxy Halos Kyle Stewart NASA Postdoctoral Fellow Jet Propulsion Laboratory, California Institute of Technology Mentor: Leonidas Moustakas The Baryon Cycle, UC Irvine,

More information

Neutron Stars. Neutron Stars and Black Holes. The Crab Pulsar. Discovery of Pulsars. The Crab Pulsar. Light curves of the Crab Pulsar.

Neutron Stars. Neutron Stars and Black Holes. The Crab Pulsar. Discovery of Pulsars. The Crab Pulsar. Light curves of the Crab Pulsar. Chapter 11: Neutron Stars and Black Holes A supernova explosion of an M > 8 M sun star blows away its outer layers. Neutron Stars The central core will collapse into a compact object of ~ a few M sun.

More information

arxiv: v1 [astro-ph.ga] 1 Dec 2016

arxiv: v1 [astro-ph.ga] 1 Dec 2016 Gas Accretion and Angular Momentum Kyle R. Stewart arxiv:1612.00513v1 [astro-ph.ga] 1 Dec 2016 Abstract In this chapter, we review the role of gas accretion to the acquisition of angular momentum, both

More information

AGN Feedback. Andrew King. Dept of Physics & Astronomy, University of Leicester Astronomical Institute, University of Amsterdam. Heidelberg, July 2014

AGN Feedback. Andrew King. Dept of Physics & Astronomy, University of Leicester Astronomical Institute, University of Amsterdam. Heidelberg, July 2014 AGN Feedback Andrew King Dept of Physics & Astronomy, University of Leicester Astronomical Institute, University of Amsterdam Heidelberg, July 2014 galaxy knows about central SBH mass velocity dispersion

More information

The Birth Of Stars. How do stars form from the interstellar medium Where does star formation take place How do we induce star formation

The Birth Of Stars. How do stars form from the interstellar medium Where does star formation take place How do we induce star formation Goals: The Birth Of Stars How do stars form from the interstellar medium Where does star formation take place How do we induce star formation Interstellar Medium Gas and dust between stars is the interstellar

More information

Chapter 21 Galaxy Evolution. How do we observe the life histories of galaxies?

Chapter 21 Galaxy Evolution. How do we observe the life histories of galaxies? Chapter 21 Galaxy Evolution How do we observe the life histories of galaxies? Deep observations show us very distant galaxies as they were much earlier in time (old light from young galaxies). 1 Observing

More information

The Star Clusters of the Magellanic Clouds

The Star Clusters of the Magellanic Clouds The Dance of Stars MODEST-14 The Star Clusters of the Magellanic Clouds Eva K. Grebel Astronomisches Rechen-Institut Zentrum für Astronomie der Universität Heidelberg Star Clusters in the Magellanic Clouds!

More information

Princeton December 2009 The fine-scale structure of dark matter halos

Princeton December 2009 The fine-scale structure of dark matter halos Princeton December 2009 The fine-scale structure of dark matter halos Simon White Max Planck Institute for Astrophysics The dark matter structure of CDM halos A rich galaxy cluster halo Springel et al

More information

ASTR 610 Theory of Galaxy Formation Lecture 18: Disk Galaxies

ASTR 610 Theory of Galaxy Formation Lecture 18: Disk Galaxies ASTR 610 Theory of Galaxy Formation Lecture 18: Disk Galaxies Frank van den Bosch Yale University, spring 2017 The Structure & Formation of Disk Galaxies In this lecture we discuss the structure and formation

More information

Quasars ASTR 2120 Sarazin. Quintuple Gravitational Lens Quasar

Quasars ASTR 2120 Sarazin. Quintuple Gravitational Lens Quasar Quasars ASTR 2120 Sarazin Quintuple Gravitational Lens Quasar Quasars Quasar = Quasi-stellar (radio) source Optical: faint, blue, star-like objects Radio: point radio sources, faint blue star-like optical

More information

Black Holes and Active Galactic Nuclei

Black Holes and Active Galactic Nuclei Black Holes and Active Galactic Nuclei A black hole is a region of spacetime from which gravity prevents anything, including light, from escaping. The theory of general relativity predicts that a sufficiently

More information

Astronomy 730. Evolution

Astronomy 730. Evolution Astronomy 730 Evolution Outline } Evolution } Formation of structure } Processes on the galaxy scale } Gravitational collapse, merging, and infall } SF, feedback and chemical enrichment } Environment }

More information

AS1001:Extra-Galactic Astronomy

AS1001:Extra-Galactic Astronomy AS1001:Extra-Galactic Astronomy Lecture 5: Dark Matter Simon Driver Theatre B spd3@st-andrews.ac.uk http://www-star.st-and.ac.uk/~spd3 Stars and Gas in Galaxies Stars form from gas in galaxy In the high-density

More information

high density low density Rayleigh-Taylor Test: High density medium starts on top of low density medium and they mix (oil+vinegar) Springel (2010)

high density low density Rayleigh-Taylor Test: High density medium starts on top of low density medium and they mix (oil+vinegar) Springel (2010) GAS MIXES high density Springel (2010) low density Rayleigh-Taylor Test: High density medium starts on top of low density medium and they mix (oil+vinegar) HOT HALO highest resolved density nth= 50x10

More information

Number of Stars: 100 billion (10 11 ) Mass : 5 x Solar masses. Size of Disk: 100,000 Light Years (30 kpc)

Number of Stars: 100 billion (10 11 ) Mass : 5 x Solar masses. Size of Disk: 100,000 Light Years (30 kpc) THE MILKY WAY GALAXY Type: Spiral galaxy composed of a highly flattened disk and a central elliptical bulge. The disk is about 100,000 light years (30kpc) in diameter. The term spiral arises from the external

More information

A new mechanism for the formation of PRGs

A new mechanism for the formation of PRGs A new mechanism for the formation of PRGs Spavone Marilena (INAF-OAC) Iodice Enrica (INAF-OAC), Arnaboldi Magda (ESO-Garching), Longo Giuseppe (Università Federico II ), Gerhard Ortwin (MPE-Garching).

More information

Galaxy Formation: Overview

Galaxy Formation: Overview Galaxy Formation: Overview Houjun Mo March 30, 2004 The basic picture Formation of dark matter halos. Gas cooling in dark matter halos Star formation in cold gas Evolution of the stellar populaion Metal

More information

arxiv:astro-ph/ v1 17 Aug 2001

arxiv:astro-ph/ v1 17 Aug 2001 HOW DID GLOBULAR CLUSTERS FORM? Sidney van den Bergh Dominion Astrophysical Observatory, Herzberg Institute of Astrophysics, National arxiv:astro-ph/0108298v1 17 Aug 2001 Research Council of Canada, 5071

More information

Two Phase Formation of Massive Galaxies

Two Phase Formation of Massive Galaxies Two Phase Formation of Massive Galaxies Focus: High Resolution Cosmological Zoom Simulation of Massive Galaxies ApJ.L.,658,710 (2007) ApJ.,697, 38 (2009) ApJ.L.,699,L178 (2009) ApJ.,725,2312 (2010) ApJ.,744,63(2012)

More information

Dwarf Galaxies as Cosmological Probes

Dwarf Galaxies as Cosmological Probes Dwarf Galaxies as Cosmological Probes Julio F. Navarro The Ursa Minor dwarf spheroidal First Light First Light The Planck Satellite The Cosmological Paradigm The Clustering of Dark Matter The Millennium

More information

Galaxy formation and evolution II. The physics of galaxy formation

Galaxy formation and evolution II. The physics of galaxy formation Galaxy formation and evolution II. The physics of galaxy formation Gabriella De Lucia Astronomical Observatory of Trieste Outline: ü Observational properties of galaxies ü Galaxies and Cosmology ü Gas

More information

The Intergalactic Medium: Overview and Selected Aspects

The Intergalactic Medium: Overview and Selected Aspects The Intergalactic Medium: Overview and Selected Aspects Draft Version Tristan Dederichs June 18, 2018 Contents 1 Introduction 2 2 The IGM at high redshifts (z > 5) 2 2.1 Early Universe and Reionization......................................

More information

Elad Zinger Hebrew University Jerusalem Spineto, 12 June Collaborators: Avishai Dekel, Yuval Birnboim, Daisuke Nagai & Andrey Kravtsov

Elad Zinger Hebrew University Jerusalem Spineto, 12 June Collaborators: Avishai Dekel, Yuval Birnboim, Daisuke Nagai & Andrey Kravtsov Elad Zinger Hebrew University Jerusalem IGM@50, Spineto, 12 June 2015 Collaborators: Avishai Dekel, Yuval Birnboim, Daisuke Nagai & Andrey Kravtsov They re still there! Account for most of the accretion.

More information

Stellar evolution Part I of III Star formation

Stellar evolution Part I of III Star formation Stellar evolution Part I of III Star formation The interstellar medium (ISM) The space between the stars is not completely empty, but filled with very dilute gas and dust, producing some of the most beautiful

More information

Dark matter and galaxy formation

Dark matter and galaxy formation Dark matter and galaxy formation Galaxy rotation The virial theorem Galaxy masses via K3 Mass-to-light ratios Rotation curves Milky Way Nearby galaxies Dark matter Baryonic or non-baryonic A problem with

More information

Summary So Far! M87van der Maerl! NGC4342! van den Bosch! rotation velocity!

Summary So Far! M87van der Maerl! NGC4342! van den Bosch! rotation velocity! Summary So Far Fundamental plane connects luminosity, scale length, surface brightness, stellar dynamics. age and chemical composition Elliptical galaxies are not randomly distributed within the 3D space

More information

Components of Galaxies: Dark Matter

Components of Galaxies: Dark Matter Components of Galaxies: Dark Matter Dark Matter: Any Form of matter whose existence is inferred solely through its gravitational effects. -B&T, pg 590 Nature of Major Component of Universe Galaxy Formation

More information

Observational Evidence of AGN Feedback

Observational Evidence of AGN Feedback 10 de maio de 2012 Sumário Introduction AGN winds Galaxy outflows From the peak to the late evolution of AGN and quasars Mergers or secular evolution? The AGN feedback The interaction process between the

More information

Lecture 7: the Local Group and nearby clusters

Lecture 7: the Local Group and nearby clusters Lecture 7: the Local Group and nearby clusters in this lecture we move up in scale, to explore typical clusters of galaxies the Local Group is an example of a not very rich cluster interesting topics include:

More information

Gas in and around z > 2 galaxies

Gas in and around z > 2 galaxies Gas in and around z > 2 galaxies Michele Fumagalli August 2010 Santa Cruz Xavier Prochaska Daniel Kasen Avishai Dekel In collaboration with: Daniel Ceverino Joel Primack Gas in galaxies from theory Gas

More information

Large Scale Structure

Large Scale Structure Large Scale Structure Measuring Distance in Universe-- a ladder of steps, building from nearby Redshift distance Redshift = z = (λ observed - λ rest )/ λ rest Every part of a distant spectrum has same

More information

A100 Exploring the Universe: The Milky Way as a Galaxy. Martin D. Weinberg UMass Astronomy

A100 Exploring the Universe: The Milky Way as a Galaxy. Martin D. Weinberg UMass Astronomy A100 Exploring the Universe: The Milky Way as a Galaxy Martin D. Weinberg UMass Astronomy astron100-mdw@courses.umass.edu November 12, 2014 Read: Chap 19 11/12/14 slide 1 Exam #2 Returned and posted tomorrow

More information

The Superbubble Power Problem: Overview and Recent Developments. S. Oey

The Superbubble Power Problem: Overview and Recent Developments. S. Oey The Superbubble Power Problem: Overview and Recent Developments S. Oey It has been known for decades that superbubbles generated by massive star winds and supernovae are smaller than expected based on

More information

GALAXIES 626. The Milky Way II. Chemical evolution:

GALAXIES 626. The Milky Way II. Chemical evolution: GALAXIES 626 The Milky Way II. Chemical evolution: Chemical evolution Observation of spiral and irregular galaxies show that the fraction of heavy elements varies with the fraction of the total mass which

More information

Structure of Dark Matter Halos

Structure of Dark Matter Halos Structure of Dark Matter Halos Dark matter halos profiles: DM only: NFW vs. Einasto Halo concentration: evolution with time Dark matter halos profiles: Effects of baryons Adiabatic contraction Cusps and

More information

Topics for Today s Class

Topics for Today s Class Foundations of Astronomy 13e Seeds Chapter 11 Formation of Stars and Structure of Stars Topics for Today s Class 1. Making Stars from the Interstellar Medium 2. Evidence of Star Formation: The Orion Nebula

More information

Galaxy Hydrodynamic Simulations and Sunrise Visualizations

Galaxy Hydrodynamic Simulations and Sunrise Visualizations Galaxy Hydrodynamic Simulations and Sunrise Visualizations Joel Primack, UCSC Daniel Ceverino, HU Madrid Avishai Dekel, HU & UCSC Sandra Faber, UCSC Anatoly Klypin, NMSU Patrik Jonsson, Harvard CfA Chris

More information

The Milky Way in the cosmological context. Andrey Kravtsov The University of Chicago

The Milky Way in the cosmological context. Andrey Kravtsov The University of Chicago The Milky Way in the cosmological context Andrey Kravtsov The University of Chicago Milky Way and Its Stars, KITP, 2 February 2015 Cosmological context: hierarchical structure formation from a Gaussian

More information

Payne-Scott workshop on Hyper Compact HII regions Sydney, September 8, 2010

Payne-Scott workshop on Hyper Compact HII regions Sydney, September 8, 2010 Payne-Scott workshop on Hyper Compact HII regions Sydney, September 8, 2010 Aim Review the characteristics of regions of ionized gas within young massive star forming regions. Will focus the discussion

More information

Age-redshift relation. The time since the big bang depends on the cosmological parameters.

Age-redshift relation. The time since the big bang depends on the cosmological parameters. Age-redshift relation The time since the big bang depends on the cosmological parameters. Lyman Break Galaxies High redshift galaxies are red or absent in blue filters because of attenuation from the neutral

More information

The Milky Way Galaxy and Interstellar Medium

The Milky Way Galaxy and Interstellar Medium The Milky Way Galaxy and Interstellar Medium Shape of the Milky Way Uniform distribution of stars in a band across the sky lead Thomas Wright, Immanuel Kant, and William Herschel in the 18th century to

More information

The cosmic distance scale

The cosmic distance scale The cosmic distance scale Distance information is often crucial to understand the physics of astrophysical objects. This requires knowing the basic properties of such an object, like its size, its environment,

More information

The visible constituents of the Universe: Non-relativistic particles ( baryons ): Relativistic particles: 1. radiation 2.

The visible constituents of the Universe: Non-relativistic particles ( baryons ): Relativistic particles: 1. radiation 2. The visible constituents of the Universe: Non-relativistic particles ( baryons ): Galaxies / Clusters / Super-clusters Intergalactic Medium Relativistic particles: 1. radiation 2. neutrinos Dark sector

More information

Astr 2310 Thurs. March 23, 2017 Today s Topics

Astr 2310 Thurs. March 23, 2017 Today s Topics Astr 2310 Thurs. March 23, 2017 Today s Topics Chapter 16: The Interstellar Medium and Star Formation Interstellar Dust and Dark Nebulae Interstellar Dust Dark Nebulae Interstellar Reddening Interstellar

More information

Cooling, dynamics and fragmentation of massive gas clouds: clues to the masses and radii of galaxies and clusters

Cooling, dynamics and fragmentation of massive gas clouds: clues to the masses and radii of galaxies and clusters of massive gas and radii of M. Rees, J. Ostriker 1977 March 5, 2009 Talk contents: The global picture The relevant theory Implications of the theory Conclusions The global picture Galaxies and have characteristic

More information

Feedback and Galaxy Formation

Feedback and Galaxy Formation Heating and Cooling in Galaxies and Clusters Garching August 2006 Feedback and Galaxy Formation Simon White Max Planck Institute for Astrophysics Cluster assembly in ΛCDM Gao et al 2004 'Concordance'

More information

Rupali Chandar University of Toledo

Rupali Chandar University of Toledo Star Formation in Clusters Rupali Chandar University of Toledo Star Clusters: Near and Far. Very Near: ONC (400 pc) ~103 Msun Near: NGC 3603 (7 kpc) ~104 Msun Star Clusters: Near and Far. Kinda Near: R136

More information

Active Galactic Nuclei-I. The paradigm

Active Galactic Nuclei-I. The paradigm Active Galactic Nuclei-I The paradigm An accretion disk around a supermassive black hole M. Almudena Prieto, July 2007, Unv. Nacional de Bogota Centers of galaxies Centers of galaxies are the most powerful

More information

arxiv:astro-ph/ v1 31 Jul 1998

arxiv:astro-ph/ v1 31 Jul 1998 Cosmic Star Formation from the Milky Way and its Satellites arxiv:astro-ph/973v1 31 Jul 199 F.D.A. Hartwick Department of Physics and Astronomy, University of Victoria, Victoria, BC, Canada, VW 3P Abstract.

More information

High Redshift Universe

High Redshift Universe High Redshift Universe Finding high z galaxies Lyman break galaxies (LBGs) Photometric redshifts Deep fields Starburst galaxies Extremely red objects (EROs) Sub-mm galaxies Lyman α systems Finding high

More information

The Milky Way Galaxy. Some thoughts. How big is it? What does it look like? How did it end up this way? What is it made up of?

The Milky Way Galaxy. Some thoughts. How big is it? What does it look like? How did it end up this way? What is it made up of? Some thoughts The Milky Way Galaxy How big is it? What does it look like? How did it end up this way? What is it made up of? Does it change 2 3 4 5 This is not a constant zoom The Milky Way Almost everything

More information

Our Galaxy. We are located in the disk of our galaxy and this is why the disk appears as a band of stars across the sky.

Our Galaxy. We are located in the disk of our galaxy and this is why the disk appears as a band of stars across the sky. Our Galaxy Our Galaxy We are located in the disk of our galaxy and this is why the disk appears as a band of stars across the sky. Early attempts to locate our solar system produced erroneous results.

More information

Secular Evolution of Galaxies

Secular Evolution of Galaxies Secular Evolution of Galaxies Outline:!Disk size evolution! Bar fraction vs mass & color! AM transfers, radial migrations! Bulges, thick disks Françoise Combes Durham, 19 July 2011 Two modes to assemble

More information

Stellar Populations in the Galaxy

Stellar Populations in the Galaxy Stellar Populations in the Galaxy Stars are fish in the sea of the galaxy, and like fish they often travel in schools. Star clusters are relatively small groupings, the true schools are stellar populations.

More information

Cinthya Herrera (NAOJ)

Cinthya Herrera (NAOJ) Cinthya Herrera (NAOJ) ASTE/ALMA Development Workshop 2014, June 18th, 2014 Galaxies interactions... Key in hierarchical model of galaxy formation and evolution (e.g., Kauffmann et al. 1993) Most massive

More information

Unstable Disks: Gas and Stars via an analytic model

Unstable Disks: Gas and Stars via an analytic model Unstable Disks: Gas and Stars via an analytic model Marcello Cacciato in collaboration with Avishai Dekel Minerva Fellow @ HUJI Theoretical studies and hydrodynamical cosmological simulations have shown

More information

A100H Exploring the Universe: Quasars, Dark Matter, Dark Energy. Martin D. Weinberg UMass Astronomy

A100H Exploring the Universe: Quasars, Dark Matter, Dark Energy. Martin D. Weinberg UMass Astronomy A100H Exploring the :, Dark Matter, Dark Energy Martin D. Weinberg UMass Astronomy astron100h-mdw@courses.umass.edu April 19, 2016 Read: Chaps 20, 21 04/19/16 slide 1 BH in Final Exam: Friday 29 Apr at

More information

Chapter 16 Lecture. The Cosmic Perspective Seventh Edition. Star Birth Pearson Education, Inc.

Chapter 16 Lecture. The Cosmic Perspective Seventh Edition. Star Birth Pearson Education, Inc. Chapter 16 Lecture The Cosmic Perspective Seventh Edition Star Birth 2014 Pearson Education, Inc. Star Birth The dust and gas between the star in our galaxy is referred to as the Interstellar medium (ISM).

More information

Disk Formation and the Angular Momentum Problem. Presented by: Michael Solway

Disk Formation and the Angular Momentum Problem. Presented by: Michael Solway Disk Formation and the Angular Momentum Problem Presented by: Michael Solway Papers 1. Vitvitska, M. et al. 2002, The origin of angular momentum in dark matter halos, ApJ 581: 799-809 2. D Onghia, E. 2008,

More information

Galaxy interaction and transformation

Galaxy interaction and transformation Galaxy interaction and transformation Houjun Mo April 13, 2004 A lot of mergers expected in hierarchical models. The main issues: The phenomena of galaxy interaction: tidal tails, mergers, starbursts When

More information

Phys/Astro 689: Lecture 8. Angular Momentum & the Cusp/Core Problem

Phys/Astro 689: Lecture 8. Angular Momentum & the Cusp/Core Problem Phys/Astro 689: Lecture 8 Angular Momentum & the Cusp/Core Problem Summary to Date We first learned how to construct the Power Spectrum with CDM+baryons. Found CDM agrees with the observed Power Spectrum

More information

High-Energy Astrophysics Lecture 6: Black holes in galaxies and the fundamentals of accretion. Overview

High-Energy Astrophysics Lecture 6: Black holes in galaxies and the fundamentals of accretion. Overview High-Energy Astrophysics Lecture 6: Black holes in galaxies and the fundamentals of accretion Robert Laing Overview Evidence for black holes in galaxies and techniques for estimating their mass Simple

More information

Gasdynamical and radiative processes, gaseous halos

Gasdynamical and radiative processes, gaseous halos Gasdynamical and radiative processes, gaseous halos Houjun Mo March 19, 2004 Since luminous objects, such as galaxies, are believed to form through the collapse of baryonic gas, it is important to understand

More information

Astronomy 422. Lecture 15: Expansion and Large Scale Structure of the Universe

Astronomy 422. Lecture 15: Expansion and Large Scale Structure of the Universe Astronomy 422 Lecture 15: Expansion and Large Scale Structure of the Universe Key concepts: Hubble Flow Clusters and Large scale structure Gravitational Lensing Sunyaev-Zeldovich Effect Expansion and age

More information

Accretion Disks. Review: Stellar Remnats. Lecture 12: Black Holes & the Milky Way A2020 Prof. Tom Megeath 2/25/10. Review: Creating Stellar Remnants

Accretion Disks. Review: Stellar Remnats. Lecture 12: Black Holes & the Milky Way A2020 Prof. Tom Megeath 2/25/10. Review: Creating Stellar Remnants Lecture 12: Black Holes & the Milky Way A2020 Prof. Tom Megeath Review: Creating Stellar Remnants Binaries may be destroyed in white dwarf supernova Binaries be converted into black holes Review: Stellar

More information